<<

Umeå University Medical Dissertations New Series No 1099- ISSN 0346 6612 ISBN 978-91-7264-273-7 From the department of Clinical Sciences, Obstetrics and Gynecology Umeå University

Effects of neuroactive on the recombinant GABAA in Xenopus oocyte

Mozibur Rahman

Umeå 2007 Department of Clinical Sciences, Obstetrics and Gynecology Umeå University SE-901 85 Umeå, Sweden

Cover illustration: The Revealed, created by Graham T. Johnson for the Howard Hughes Medical Institute Bulletin

Copyright© 2007 by Mozibur Rahman ISSN 0346-6612 ISBN 978-91-7264-273-7 Printed by Print & Media, Umeå University, 2007

2

To my family and parents

3

Contents

page Abstract 7 List of original papers 8 Abbreviation 9 Introduction 10 1. Relationship between the neuroactive metabolites of sex and stress hormones and disorders 10 2. of naturally occurring neuroactive steroids acting on GABAA receptors 12 2.1 Synthesis of 3α-OH steroids 13 2.2 Synthesis of 3β-OH steroids and 13 2.3 Synthesis of 5α-androstane 3α,17β-diol 14 2.4 Synthesis of THDOC 14 3. Concentration of 3α/3β neuroactive steroids in plasma and brain 16 4. GABAA receptor 16 4.1 GABAergic system 16 4.2 GABAA receptor subunit combinations and their proportion 17 4.3 Distribution of GABAA receptor in the brain 18 4.4 Pharmacological property of GABAA receptor 19 4.4.1 Allsosteric effect 19 4.4.2 Partial agonist and ‘superagonists’ of GABAA receptor 19 4.4.3 20 4.5. and GABAA receptor 20 2+ 4.6 Zn and GABAA receptor 21 4.7. Mechanism of GABAA receptor kinetics (deactivation desensitization, synaptic and extrasynaptic, phasic and tonic, duration of channel opening) 21 4.8 Pharmacological effects depend on GABA receptor subunit composition 23 4.9 GABAA receptor and disorders 24 5. Neuroactive agonists 25 5.1 Synaptic effect of agonist neuroactive steroids 25 5.2 Extrasynaptic effect of agonist neuroactive steroids 27 5.3 Subunit dependence of neuroactive steroids 28

4 5.4 Structure activity relationship 29 5.4.1 Backbone structure of neuroactive steroids 29 5.4.2 Steroid- Mimetic Core 30 5.4.3 Replacing C5-position 30 5.4.4 Replacing C11-position 30 5.4.5 Replacing C20-position 31 5.4.6 Replacing C21-position 31 5.4.7 Steroid nucleus 31 5.5 Binding sites of neuroactive steroids 31 5.6 Physiological role of neuroactive steroids 32 5.7 Neuroactive steroids as potential 33 5.7.1 effect of neuroactive steroids 33 5.7.2 Effect of neuroactive steroids in learning and memory 34 5.7.3 Effect of neuroactive steroids on sleep 34 5.7.4 Effect as 35 5.8.5 Analgesic effect of neuroactive steroids 35 6. Neuroactive steroid antagonists 35 6.1 and other sulfated steroids 35 6.1.1 Modulation of GABAA receptor by pregnenolone sulfate 35 6.1.2 Mechanism of action of pregnenolone sulfate 36 6.1.3 Site of action of pregnenolone sulfate 37 6.2 3β-hydroxy steroids 37 6.2.1 Mode of action of 3β-hydroxy steroids 37 6.2.2 Modification of 3β-position 38 6.2.3 Mechanism of action of 3β-hydroxy steroids 39 7. Recombinant GABA receptor: Advantages and disadvantages with studies on recombinant receptors expressed in oocyte over cell lines 40 8. Uses of rodent models for the testing of drugs for preclinical evaluation 41 Aims 42 Method 43 Chemicals 43 cDNA cloning 43 In vitro transcription and expression in the Xenopus oocyte 43 Oocyte electrophysiology 44 Data analysis 45 Statistical analysis 45 Results 46 GABA site antagonism and inhibition by is

5 not dependant on α-subunit (Paper I) 46 Pregnenolone sulfate response in not dependant on γ-subunit (Paper II) 46 Pregnenolone sulfate and Zn2+ do not share the same binding site (Paper II) 46 Functional differences between 3β-hydroxysteroids and pregnenolone sulfate (Paper III) 47 The neuroactive steroids activity is different between the recombinant human and rat α1β2γ2L GABAA receptors (Paper IV) 47 Neuroactive steroids activity differs between 2L (long form) and 2S (short from) variant of γ-subunit of human GABAA receptor (Paper IV) 48 Discussion 48 GABA and pentobarbital inhibition by bicuculline at α1-, α4-, 5- subunit containing GABAA receptor 48 Pregnenolone sulfate response is not dependant on γ-subunit 49 3β-hydroxysteroids and pregnenolone sulfate inhibit recombinant rat GABAA receptor through different mechanisms 50 The neuroactive steroids activity between the human and rat α1β2γ2L recombinant GABAA receptors 50 Differential response of neuroactive steroids activity between 2L (long form) and 2S (short from) variant of γ-subunit of human GABAA receptor 51 Conclusions 52 Acknowledgements 53 References 55 Appendix Paper I - IV

6 Abstract Introduction: Neuroactive steroids represent a class of both synthetic and naturally occurring steroids that have an effect on neural function. In addition to classical genomic mechanism by the hormones , deoxycorticosterone and , the 3α- OH metabolites of these hormones enhance GABAA receptor through rapid non-genomic mechanism. The site(s) of action of these neuroactive steroids namely 3α-OH-5α-pregnan- 20 one (3α5αP), (3α,5α)-3,21-deoxycorticosterone(3α5α-THDOC) and 5α androstane- 3α,17β-diol on GABAA receptor are distinct from that of and binding sites. The modulation site(s) has a well-defined structure activity relationship with a 3α-hydroxy and a 20-ketone configuration in the pregnane molecule required for agonistic action. Pregnenolone sulfate is a noncompetitive GABAA and inhibit GABA activated Cl- current in an activation dependant manner. 3β-hydroxy A-ring reduced pregnane steroids are also GABAA receptor antagonist and inhibit GABAA receptor function and its potentiation induced by their 3α-diesteromers in a noncompetitive manner. Aim: The aim was to investigate if the effect of GABA, pentobarbital antagonism by bicuculline and if the effect of GABA-agonist and antagonist neuroactive steroids including pregnenolone sulfate is dependant on the α-subunits of GABAA receptor. Furthermore, the studies aimed at investigating the binding site of pregnenolone sulfate and if its effect is dependent on γ-subunit. In addition, the inhibitory effect of pregnenolone sulfate and 3β- hydroxy steroids has been characterized. We also wanted to investigate if the neuroactive steroids effect vary between the human and rat recombinant α1β2γ2L receptors and between the long (L) and short (S) variants of γ2-subunit. Method: Experiments were performed by the two electrodes voltage-clamp technique using oocytes of Xenopus laevis expressed with recombinant GABAA receptors containing α1, α4 or α5, β2, γ2L and γ2S-subunits. Results: There was no difference between the α1, α4 and α5-containing subunits regarding GABA and pentobarbital inhibition by bicuculline. GABA-activated current in the binary αβ was potent than that of ternary αβγ receptor. Unlike Zn2+ effect, inhibition by pregnenolone sulfate on the GABAA receptor is not dependant on the γ-subunit. It is likely that the 2’ residue closest to the N-terminus of the protein at M2 helix on both α1 and β2 subunit are critical to the inhibitory actions of PS and the function of Cl- channels. Point mutation at M2 helix of the β2-subunit (β2A252S) can dramatically reduce the inhibitory effect of PS on the GABAA receptors without affecting the inhibitory properties of 3β- hydroxysteroids. Agonist and antagonist steroids also varied in their efficacy between the human and rat α1β2γ2L receptor. Neuroactive steroids also showed difference between human γ2L and γ2S-containing receptor. Conclusions: GABA and pentobarbital antagonism by bicuculline is not dependant on α- subunit. Pregnenolone sulfate binding site is different from that of Zn2+. 3β- hydroxysteroids and pregnenolone sulfate inhibit GABAA receptor through different mechanisms. Neuroactive steroids also differ between species and between the long and short variant of γ- subunit. Key words GABA, GABAA receptor, neuroactive steroids

7 Lists of original papers*

Paper I. GABA-site antagonism and pentobarbital actions do not depend on the α-subunit type in the recombinant rat GABA-A receptor. Rahman, M., Zhu, D., Lindblad, C., Johansson, I., Holmberg, E., Isaksson, M., Taube, M., Bäckström, T., Wang, M., 2006. Acta physiol 187, 479-488.

Paper II. Pregnenolone sulfate and inhibit recombinant rat GABAA receptor through different channel property. Wang, M., Rahman, M., Zhu, D., Bäckström, T., 2006. Acta physiol 188, 153 - 163.

Paper III. 3β-hydroxysteroids and pregnenolone sulfate inhibit recombinant rat GABAA receptor through different channel property. Ming-De Wang, Mozibur Rahman, Di Zhu, Inga-Maj Johansson and Torbjörn Bäckström. European journal of 557, 124 -131.

Paper IV. Functional difference between recombinant human and rat α1β2γ2L GABAA receptors. Mozibur Rahman, Monica Isaksson, Inga-Maj Johansson, Gianna Ragagnin, Torbjörn Bäckström and Ming-De Wang, 2007. Manuscript.

* All published papers are reproduced in this thesis with the kind permission of the copyright holders

8 Abbreviations GABA γ-amino DBI binding inhibitor 3α5αP 3α, 5α-tetrahydroprogesterone, 3α-OH-5α-pregnan-20-one, 3α5βP 3α-OH-5β-pregnan-20-one, 3α5αTHDOC 3α, 5α-tetrahydrodeoxy- 3α5βTHDOC 3α, 5β-tetrahydrodeoxy-corticosterone NMDA N-methyl-D-asparate PMDD premenstrual dystrophic disorder P450scc P-450 side chain cleavage 3α-HSD 3α-hydroxysteroid dehydrogenase 3β-HSD 3β-hydroxysteroid dehydrogenase 17β-HSD 17β-hydroxysteroid dehydrogenase 5α-R 5α-reductase 11β-HSD 11β-hydroxysteroid dehydrogenase 11βOHase 11β-hydroxylase 8-HSOR 18-hydroxysteroid oxidoreductase 21OHase 21-hydroxylase StAR steroidogenic acute-regulatory 3α-OH 3α-hydroxyl 3β-OH 3β-hydroxyl PS pregnenolone sulfate DHEAS sulfate 3β5αP/UC1010 3β-OH-5α -pregnan-20-one, Isoallopregnanolone 3β5βP/UC1014 3β-OH-5β -pregnan-20-one, UC1011 5α-pregnan-3β, 20β-diol UC1013 5β-pregnan-3β, 20α-diol UC1015 5β-pregnan-3β, 21-diol-20-one; 3β5βTHDOC UC1019 5α-pregnan-3β, 20α-diol UC2020 5β-pregnan-3β, 20β-diol 3α-adiol/3α5αADL 5α-androstane-3α,17β-diol DOC deoxycorticosterone 3α5αACN 3α-5α- androstane carbonitrile DMSO dimethyl sulfoxide ANOVA analysis of variance BMI bicuculline methiodide IPSC inhibitory post-synaptic current sIPSC spontaneous inhibitory post-synaptic current PKC protein kinase C DGC dentate gyrus granule cells CGC cerebellar granule cells

Imax the maximum current amplitude

EC50 concentration of that produces 50% of Imax

9

Introduction

1. Relationship between the neuroactive metabolites of sex and stress hormones and disorders The sex steroid hormones, and progesterone, influence a variety of behaviors in vertebrate and have a crucial effects on female reproductive function (Whitfield et al., 1999). Estrogen is required for the development of female phenotype, sexual maturation, female genital function and skeletal maintenance. Progesterone is necessary for conception and the maintenance of pregnancy. In adult women, the main sources of are the granulose cells of developing follicle and the corpus luteum in the ovary. Progesterone is synthesized mainly in the granulose cells of corpus luteum, also in the placenta and the adrenal gland (Speroff, 2005). Two progesterone receptors (PR) are also known, PRA and PRB (Alves et al., 1998; Bethea, 1993). These hormones are known to act through genomic mechanism in which intracellular receptor are located in the nucleus or cytoplasm and act as -activated transcription factors in the regulation of gene expression (Fig. 1). However, metabolites of progesterone and several stress hormones have effects through membrane bound receptors such as ion channels; a fast and non-genomic mechanism in which receptor binding to DNA and RNA synthesis is not required (Fig. 1) (Baulieu and Robel, 1995; Frye et al., 1992; Rupprecht, 2003). While the action of steroids at the genome requires a time period from minutes to hours that is limited by the rate of protein biosynthesis (McEwen, 1991), the modulatory effects of metabolites are fast occurring events requiring only milliseconds to seconds (McEwen, 1991). There is increasing evidence that these metabolites which act via non-genomic pathways may alter neuronal excitability (Lambert et al., 1995; Majewska et al., 1986; Paul and Purdy, 1992). The term ‘neuroactive steroids’ has been coined for these metabolites with this particular property (Paul and Purdy, 1992). The 3α-reduced metabolites of progesterone and deoxycorticosterone namely, 3α, 5α-tetrahydroprogesterone (3α5αP; allopregnanolone) and 3α, 5α-tetrahydrodeoxy-corticosterone (3α5αTHDOC) were the first steroids that have been shown to modulate neuronal excitability via their interaction with γ-aminobutyric acid type A (GABAA) receptors (Majewska et al., 1986). Several other like glutamate (N-methyl-D-asparate, NMDA) (Wu et al., 1991), nicotinic (Valera et al., 1992), muscarinic (Klangkalya and Chan, 1988), (Parry, 2001) and the system (Halbreich, 2003) are also modulated by such steroids (Fig. 1). Acting on the GABAA receptor the and metabolites may therefore exert their clinically significant effects in certain women; at low concentration, moderate to severe adverse emotional reactions are induced in up to 20% individual (Beauchamp et al., 2000; Fish et al., 2001), namely negative mood in

10 premenstrual dystrophic disorder (PMDD) (Backstrom et al., 2003; Sundstrom Poromaa et al., 2003) and petit mal (Banerjee and Snead, 1998; Grunewald et al., 1992), catamenial epilepsy -the epileptic in

Figure 1. Genomic and non-genomic site of action of steroid hormones and their metabolites.

women related to plasma estrogen and progesterone during the menstrual cycle (Backstrom, 1976). At higher doses they affect learning (Johansson et al., 2002), act as , anti-aggressive, /anesthetic and anti-epileptic effect in both animal and human (Backstrom et al., 1990; Bjorn et al., 2002; Paul and

11 Purdy, 1992; Wang et al., 2001). The main focus of this thesis is the effect of these steroid metabolites (neuroactive steroids) on the GABA-system.

2. Biosynthesis of naturally occurring neuroactive steroids acting on GABAA receptors Neuroactive steroids are formed de novo in and glia, or generated by metabolism of circulating precursors that originate in peripheral steroidogenic

Figure 2. The pathway for synthesis of allopregnanolone and pregnenolone sulfate within the or glial cell. Steroidogenic acute-regulatory (StAR) protein might interact with the mitochondrial receptor (MBR) to facilitate the transport of across the mitochondrial membrane. Abbreviations: 3α-HSD, 3α-hydroxysteroid dehydrogenase; 3β- HSD, 3β-hydroxysteroid dehydrogenase; 5α-DHP, 5α-dihydroprogesterone; P450scc, P450 side-chain cleavage. (Adapted from Belelli and Lambert, 2005).

12 organs (Compagnone and Mellon, 2000). The neuroactive steroids are A-ring reduced metabolites of the steroid hormones progesterone, deoxycorticosterone, and testosterone (Compagnone and Mellon, 2000). Their synthesis in the brain is controlled by the endogenous peptide-diazepam binding inhibitor (DBI), a ligand for the ‘peripheral’ benzodiazepine binding site, which are independent of GABA binding sites (Costa et al., 1994). Synthesis of different groups of neuroactive steroids is described in the next paragraph. Brain astrocytes and neuron express cytochrome P-450 side chain cleavage cytochrome (P450scc) which converts cholesterol to pregnenolone an immediately necessary for the synthesis of all hormonal steroids (Fig. 2) (Patte-Mensah et al., 2003; Zwain and Yen, 1999). Steroidogenic acute-regulatory protein (StAR) also facilitates the conversion. 3β-hydrxysteroid dehydrogenase (3β-HSD) converts pregnenolone to progesterone. Other that mediate the synthesis of neuroactive steroids are cytochrome P-450c17 (P450c17); 3α-hydroxysteroid dehydrogenase (3α-HSD); 17β hydroxysteroid dehydrogenase(17β-HSD); 5α-reductase(5α-R); 5α- (5α-DHT); 11β-hydroxysteroid dehydrogenase(11β-HSD); Sulfotransferase and sulfatase;11β-hydroxylase (11βOHase); 18-hydroxylase (18βOHase); 18-hydroxysteroid oxidoreductase (18-HSOR); 21- hydroxylase(21OHase) (Mensah-Nyagan et al., 1999).

2.1 Synthesis of 3α-OH steroids For the synthesis of allopregnanolone (3α-OH-5α-pregnan-20-one) from progesterone, 5α-reductase and 3α-hydroxysteriod dehydrogenase are needed, whereas its 5β-steroid-isomer, pregnanolone (3α-OH-5β-pregnan-20-one) is produced by the enzymatic activity of 5β-reductase and 3α-hydroxysteroid dehydrogenase (3α-HSD) (Fig. 2). Because the activity of the 3α-HSD is far greater than that of the 5α-reductase, steroid 5α-reductase is the rate-limiting steps in the biosynthesis of the neuroactive steroids (Penning et al., 1985; Russell and Wilson, 1994). Allopregnanolone and pregnenolone have high affinity to the GABAA receptor (Majewska et al., 1986). Allopregnanolone has been found to be the most potent of the progesterone metabolite, followed by pregnenolone (Paul and Purdy, 1992), (Zhu et al., 2001). Allopregnanolone persist in the brain after adrenalectomy and gonadectomy or after pharmacological suppression of adrenal and gonadal suppression (Cheney et al., 1995; Corpechot et al., 1993; Purdy et al., 1991), indicating that this steroid can be synthesized de novo in the brain via 5α- reduction of progesterone.

2.2 Synthesis of 3β-OH steroids and pregnenolone sulfate Pregnenolone sulfate (PS) and dehydroepiandrosterone sulfate (DHEAS) are also endogenously produced GABA active steroids (Paul and Purdy, 1992).

13 Pregnenolone sulfate is synthesized from pregnenolone and mediated by the sulfotransferase (Fig. 2). Conversion from dehydroepiandrosterone to dehydroepiandrosterone sulfate (DHEAS) is also mediated by sulfotransferase. Dehydroepiantrosterone is metabolized from pregnenolone by cytochrome P450C17. Epipregnanolone (3β-OH-5β -pregnan-20-one) and other 3β-OH are also evident to be synthesized in the body (Hill et al., 2001). They are synthesized from O O O

5α−reductase 3β-HSD

O O HO H H β α Progesterone 5α-DHP 3 5 P (Isoallopregnanolone) Figure 3. Synthesis of 3β-OH steroids (Isopregnanolone) progesterone. 5α or 5β-dihdroprogesterone is synthesized from progesterone by enzyme 5α-or 5β-reductase. 3β-hydroxysteroid dehydrogenase (3β-HSD) is essential for the conversion to 3β-hydroxy steroids. Isoallopregnanolone (3β-OH- 5α -pregnan-20-one) is produced in the same way to allopregnanolone with the exception that 3β-HSD is needed in stead of 3α-HSD (Fig. 3).

2.3 Synthesis of 5α-androstane 3α, 17β-diol Synthesis of potent dihydrotestosterone from the precursor testosterone is mediated by steroid 5α-reductase. Breakdown to the inactive androgen metabolite 5α- androstane-3α,17β-diol (3α-adiol) is mediated by reductive 3α-HSD OH OH OH

5α-reductase 3α-HDS

RoDH1 O OH O H H Testosterone Dihydrotestosterone 5α-Androstane-3α,17β-diol Figure 4. Synthesis of 5α-androstane 3α,17β-diol. RoDH1 represents the rat retinol dehydrogenase 1 enzyme.

(Fig. 4) (Biswas and Russell, 1997). 5α-androstane-3α,17β-diol (3α-adiol) also modulate the GABA-mediated Cl- flux (Frye et al., 1996)

2.4 Synthesis of THDOC Stress hormone (3α,5α)-3,21-deoxycorticosterone(3α5α-THDOC) and (3α,5β)- 3,21-deoxycorticosterone(3α5βTHDOC) are also potent naturally occurring steroids which acts on the GABAA receptor (Crawley et al., 1986; Gasior et al., 1999; Lambert et al., 2001; Majewska et al., 1986). 3α5αTHDOC is a metabolite of

14 the deoxycorticosterone and is responsible for the sedative and anti-seizure activity of deoxycorticosterone in animal models (Reddy and Rogawski, 2002). Deoxycorticosterone (DOC) can also be metabolized from progesterone and this conversion is mediated by steroid 21 hydroxylase (P45021)(Edwards et al., 2005). The neuroactive steroid 3α5α- tetrahydrodeoxycorticosterone (3α5αTHDOC) is synthesized from deoxycorticosterone by two sequential A-ring reductions. 5α-Reductase first converts deoxycorticosterone to the intermediate 5α-dihydrodeoxycorticosterone, OH OH OH

O O O O α− 5 3α-HSD P45021 reductase O O O HO H H Progesterone Deoxycorticosterone 5α-DHDOC 3α5αTHDOC Figure 5. Synthesis of 3α5αTHDOC.

which is then further reduced by 3α-hydroxysteroid oxidoreductase to form 3α5αTHDOC (Fig. 5). 3α5αTHDOC has significant sedative effects acting on GABAA receptors. Plasma concentrations of 3α5αTHDOC (half-life <20 min) increase rapidly following systemic administration of deoxycorticosterone. In contrast to the progesterone-derived neuroactive steroid allopregnanolone, which

O O CN O O

HO HO HO H H HO H 3α5αACN alphaxalone Δ4 pregnen- 3α ol-20 one UC1010

OH HO O HO HO

H H H

HO HO HO HO H H H H UC 1011 UC1015 UC1019 UC1020

Figure 6. Other agonist and antagonist steroids tested in this project (chemical names are given in the method part)

15 is synthesized de novo in the brain, 3α5αTHDOC appears to be derived nearly exclusively from the adrenal cortex. The conversion of deoxycorticosterone into 3α5αTHDOC occurs both in peripheral tissues and in the brain (Reddy, 2003).

3. Concentration of 3α/3β neuroactive steroids in human plasma and brain (mean ± SEM) Steroids Plasma(nM) Brain(nM) Follicular Luteal Postmeno Luteal -pausal progesterone 5.0 ± 0.50a 34.7 ± 2.40 65d 137d 3α-OH-5β-pregnan-20-one 0.6 ± 0.00b 1.1 ± 0.50 - 114d 3α-OH-5α-pregnan-20-one 0.53 ±0.19c 2.14 ± 0.37 47d 66d 3β-OH-5α-pregnan-20-one 0.29 ± 0.14c 1.23 ± 0.28 - - 3β-OH-5β-pregnan-20-one 0.09 ± 0.08c 0.26 ± 0.13 - - pregnenolone sulfate 11.2 ± 0.6 a 15.2 ± 0.8 a - 100f pregnenolone *2.19g - - - 3α5α-androstane-3α,17β-diol *0.475g - - - Data are cited from following references: a (Wang et al., 1996); b(Sundstrom et al., 1998); c (Havlikova et al., 2006); d(Bixo et al., 1997); e (Bixo et al., 1995); f(Lanthier and Patwardhan, 1986) ; g(Kancheva et al., 2007).* concentration in adult men.

To our knowledge there is no report on the plasma concentration of 3α5αTHDOC in human. Although Hill et al. (2007) measured the plasma level of 3α5αTHDOC (0.3 nM/L) in adult men (unpublished, personal contact). However, in rat the plasma level normally fluctuate between 1 and 5 nM but increase to 15- 40 nM following acute stress (Reddy and Rogawski, 2002) and can reach 40-60 nM in pregnancy (Concas et al., 1998). Purdy et al. first observed that both circulating and brain level of 3α5αTHDOC reach their peak during stress; in the rat this amounts to a circulating concentration of ~20 nM and a brain tissue concentration of ~120 nM (40 ng/g) (Purdy et al., 1991). The plasma level of level in human is around 0.8- 1.2 µM/L. In stressful situation like PhD examination the cortisol level increased up to ~1.3 - 1.8 µM/L (Droogleever Fortuyn et al., 2004).

4. GABAA receptor 4.1 GABAergic system As mentioned earlier that GABA mediates most inhibitory transmission events in the vertebrate brain. GABA is synthesized by two isoforms of decarboxylase (GAD) (Erlander et al., 1991; Esclapez et al., 1994; Soghomonian and Martin, 1998). Inhibition mediated by GABA is involved crucially in both

16 short- and long-term regulation of neuronal excitability. Three different types, GABAA, GABAB and GABAC, of GABA receptor might be activated in the CNS (Sivilotti 1993). It has been estimated that approximately 33% of the in the mammalian are GABAergic (Purvez et al., 2004). GABAA receptors are members of the ligand gated ion channel superfamily (Schofield et + 2+ al., 1987). Metabotropic GABAB receptor is coupled to certain K and Ca channels via GTP binding protein (G-protein) and/or other messengers (Hill and Bowery, 1981). GABAA receptor is bicuculline sensitive and GABAB is beclofen sensitive. The third type of GABA receptor, insensitive to both bicuculline and beclofen was designated GABAC (Drew et al., 1984). The GABAC responses are also of the fast type associated with the opening of an anion channel; they are, unaffected by typical modulators of GABAA receptor like benzodiazepines and barbiturates (Bormann and Feigenspan, 1995; Sivilotti and Nistri, 1991).

4.2 GABAA receptor subunit combinations and their proportion To date, 16 isoforms of GABAA receptor have been identified (Whiting, 2003). These comprise α1-6, β1-3, γ1-3, δ, ε, π and θ. GABAA receptors are pentameric proteins (Nayeem et al., 1994) of five different subunits containing 2α/2β/1γ- or δ-subunit variants(Fig. 7) (Farrar et al., 1999; Klausberger et al., 2001; Knight et al., 2000). Immunological, pharmacological, and functional analysis give the evidence that the α1β2γ2 combination represents the largest population of GABAA receptor (~60%) followed by α2β3γ2 (~15-20%) and α3βnγ2 (~10-15%,

Figure 7. GABAA receptor with binding sites of neuroactive steroids, barbiturate, benzodiazepine and . Right figure shows the transmembrane domains of five subunits. Note that the second transmembrane domain of each subunit forms the channel lining.

17 n=1,2 or 3) (Fig. 8) (Fritschy and Brunig, 2003; Mohler et al., 2004; Mohler et al., 2002; Wallner et al., 2003). Receptors containing the α4-, α5-, and α6-suunits, as well as the β1-,γ1-γ3, δ-,π- and θ-subunits, form minor receptor population. α4βnδ, α4βnγ and α6β2,3γ2 receptors- each of these is less than 5% of all GABAA receptors (Fig. 8). α6βnδ has a small population only in the cerebellum. Likewise, α6β2,3γ2 receptors are located only in the cerebellum. The ρ-subunits are expressed primarily in the retina and correspond to the so-called GABAC receptor (Bormann, 2000). Structurally, this receptor differs from classical GABAA receptors in that the channel comprises a homopentamer of ρ-subunit rather than heteropentamers.

4.3 Distribution of GABAA receptor in the brain Expression of GABAA receptor subtypes in the adult brain exhibits a remarkable region- and neuron-specificity which suggests that individual subtypes are present in distinct neuronal circuits. α1β2γ2 receptor present in most brain areas and it is localized to interneuron in the and cortex (layer I-IV), and cerebral Purkinje cells (McKernan and Whiting, 1996). α2β3γ2 receptor present in cerebral cortex (layer I-IV), hippocampal formation, amygdale, striatum, olfactory bulb, hypothalamus, superior colliculi and motor nuclei (Fritschy and Brunig, 2003). α3βnγ2, α3γ2, α3θ are abundant in the cerebral cortex (layers V-VI), amygdala, olfactory bulb, thalamic reticular and intralaminar nuclei, superior colliculus, brainstem, spinal cord, locus coeruleus etc. α4βxδ in the dentate gyrus and thalamus. α5β3γ2 receptor is localized in hippocampal pyramidal cells, deep

Figure 8. Proportion of GABAA receptor in the central nervous system (CNS) (based on the report by Mohler et al., 2004 and Fritschy et al., 2003)

18 cortical layers, amygdala, olfactory bulb, hypothalamus, superior colliculus, superior olivary nucleus, spinal trigeminal nucleus and spinal cord. α6β2,3γ2, α6β2,3δ and α6β2,3γ2 are found mainly in the cerebellum, dorsal cochlear nucleus (Fritschy and Brunig, 2003). Other minor combinations are distributed throughout the brain.

4.4 Pharmacological property of GABAA receptor 4.4.1 Allsosteric effect

The GABAA receptor can be modulated by a number of therapeutic agents, including benzodiazepines (Macdonald and Olsen, 1994; Sieghart, 1992), barbiturates (Smith and Riskin, 1991), anaesthetics, (Harris et al., 1995), Zinc (Smart, 1992) and neuroactive steroids (Hawkinson et al., 1994; Puia et al., 1990). Pharmacological analysis of GABAA receptors revealed that α-subunits govern GABA affinity (Smith et al., 2001), α and γ subunits regulate benzodiazepine site pharmacology (Buhr et al., 1996; McKernan et al., 1995; Smith et al., 2001; Wingrove et al., 1997). In fact, the binding pocket of benzodiazepines is located between the α and γ subunit (Sigel, 2002). The α1-, α2-, α3-, and α5- GABAA receptors are diazepam-sensitive receptors, whereas the α4- and α6- GABAA receptors are insensitive to diazepam (Benson et al., 1998; Wingrove et al., 2002). α1-, α2-, α3-, and α5 receptors are distinguished further by their affinity to (α1 > α2 = α3 >> γ5) and various β-carbolines (α1 > α2= α3) (Mohler et al., 1996; Sieghart, 1995). The action of anticonvulsant loreclazole and intravenous anesthetics are also potentiators of GABAA receptors (Hill- Venning et al., 1997; Wingrove et al., 1994). Neuroactive steroids are also positive allosteric modulators of recombinant and native GABAA receptors (Lambert et al., 2001) and will be described in detail in the next section. Furthermore, novel ligands like 3-Heteroaryl-2-pyridones and ROD compounds are being introduced, which exhibit intrinsic activity only on specific GABAA receptor subtypes (Collins et al., 2002; Sigel et al., 2001). Neurochemical, electrophysiological, and behavioral evidence accumulated over the past two decades suggests that GABAA receptors also mediate certain acute and chronic actions of ethanol (Deitrich et al., 1989; Harris et al., 1998; Ueno et al., 2001). Similar to the GABAA receptors modulators mentioned above, ethanol exhibits an array of central actions, including anxiolytic, anticonvulsant, sedative-, , and general anesthetic effects, in a dose-dependent manner (Deitrich et al., 1989; Frye et al., 1981). However, some recent attempts have failed to reproduce all these findings (Borghese et al., 2006; Carta et al., 2004).

4.4.2 Partial agonist and ‘superagonists’ of GABAA receptor and THIP (4,5,6,7-tetrahydroisoxazole(5,4-c) pyridine-2-ol) are widely used as selective GABAA receptor agonist (Chebib and Johnston, 2000). THIP

19 (commonly named ), a conformationally restricted analogue of muscimol, is a potent partial agonist of high efficacy (Krogsgaard-Larsen et al., 1977). It has no effect on synaptic currents at concentration up to 3 µM in α1- containing receptor; a full agonist at α2-containig receptor whereas in the extrasynaptic receptors it is potent, high efficacious ‘super-agonist’ (Ebert et al., 2002; Ebert et al., 1994; Wafford and Ebert, 2006). Another GABA agonist 4- PIOL(5-(4-piperidyl)isoxazol-3-ol), is a low efficacy partial agonist (Ebert et al., 2002). On the other hand, Compound 5b (N, N’-1, 4-butanediylbis[4- aminobutanamide]) also produce a maximum response that was 150% that of GABA and was described as ‘super-agonist’ (Carlier et al., 2002).

4.4.3 Channel blocker Bicuculline, Zn2+, TBPS, picrotoxin and pregnenolone sulfate are the antagonists of GABAA receptor. Bicuculline is the competitive blocker of GABA. In other word it is GABA site antagonist (Krishek et al., 1996; Polenzani et al., 1991). Picrotoxin, Zn2+ and pregnenolone sulfate etc exert their effect directly binding in the ion channel (Eisenman et al., 2003; Hosie et al., 2003; Krishek et al., 1996). Pregnenolone sulfate is considered as the open Cl- channel blocker, whereas Zn2+ stabilizes the closed Cl- channel and has no effect on open Cl- channel. Picrotoxin causes a decrease in mean channel open time. It works by preferentially shifting opening channels to the briefest open state (1 msec). Experimental like and the cage convulsant t-butyl bicyclophosphorothionate (TBPS) also act to block the Cl- channel (Delorey and Olsen 1994 -Basic Neurochemistry, Chapter 18).

4.5. Barbiturates and GABAA receptor Electrophysiological and neurochemical studies have shown that general anesthetic agents can have three mechanism of action, namely (i) a potentiation of the GABA response (Evans, 1979; Study and Barker, 1981), (ii) a direct activation of GABAA receptor (Franks and Lieb, 1994; Robertson, 1989) and (iii) at high concentrations, a block of the GABA chloride channel (Robertson, 1989; Schwartz et al., 1986). The affinities and efficacies of potentiating and direct effect of pentobarbital vary with GABAA receptor subtypes. Potentiating effect over GABA EC20 is higher in the α6-subunit containing receptor compare to α1- subunit containing receptor (Thompson et al., 1996). α6-subunit also shows higher affinity and efficacy for direct effect of pentobarbital. Like α-subunits, β-subunits are also known to contribute to the effects of barbiturates on GABA receptor (Belelli et al., 1999; Thompson et al., 1996; Wafford et al., 1996). In general, the β2- and β3- subunits confer greater sensitivity and efficacy to pentobarbital than the β1-subunit for both its allosteric and agonist actions. Direct effect of pentobarbital was blocked by picrotoxin but not by competitive antagonist

20 bicuculline or SR95531, indicating that direct agonist activity of pentobarbitone is not mediated via the GABA binding site (Thompson et al., 1996). Many studies have been performed for searching the site of action of barbiturates. A number of residues within the first and second transmembrane domains (TMs) of the β-subunit have been identified as sensitive to positive modulation and/or direct activation by several anesthetics, including the barbiturates (Birnir et al., 1997; Carlson et al., 2000; Cestari et al., 2000; Chang et al., 2003; Dalziel et al., 1999; Pistis et al., 1999; Serafini et al., 2000). Another report showed that a single amino acid residue located in the N-terminal domain of the α-subunit, regulates pentobarbital efficacy (Drafts and Fisher, 2006).

2+ 4.6 Zn and GABAA receptor The transition metal ion Zn2+ is concentrated into zinc-containing neurons in the CNS and is thought to be involved in modulating the sensitivity of GABA and glutamate receptors to their respective neurotransmitters (Frederickson, 1989; Frederickson and Bush, 2001). In some neurons, the sensitivity of GABAA receptors to Zn2+ varies throughout development (Brooks-Kayal et al., 2001; Smart, 1992; Taketo and Yoshioka, 2000), and Zn2+ could be critically involved in temporal-lobe epilepsy, in which GABAA receptor subunit expression may be altered (Buhl et al., 1996; Gibbs et al., 1997; Molnar and Nadler, 2001). Zn2+ ions inhibit GABAA receptor function by an allosteric mechanism that is critically dependent on the receptor subunit composition: αβ-subunit combinations show the highest sensitivity, and αβγ-isoforms are the least sensitive, whereas αβδ subunits and αβε receptor showing intermediate sensitivities to Zn2+(Krishek et al., 1998; Saxena and Macdonald, 1996). Zn2+ antagonism is also affected by α- subunit (Draguhn et al., 1990; Smart et al., 1991; White and Gurley, 1995). Inhibitory effect of Zn2+ is reduced by exchanging α1-subunit for α2- or α3- subunits. Recently, the site of action of Zn2+ has also been elucidated. Three discrete sites that mediate Zn2+ inhibition have been identified. One is located within the ion channel, and the other two are on the external amino (N)-terminal face of the receptor at the interfaces between α- and β-subunits (Hosie et al., 2+ 2003). Low Zn sensitivity of GABAA receptors containing the γ2-subunit results from disruption to two of the three sites after receptor subunit co- assembly.

4.7. Mechanism of GABAA receptor kinetics (deactivation desensitization, synaptic and extra-synaptic, phasic and tonic, duration of channel opening GABAA receptor, like other receptors in the brain may be synaptic or extrasynaptic depending on the location of the receptors. The receptors within the synaptic cleft (bouton) are known as ‘synaptic’ whereas the receptors that are over the neuronal surface outside the synaptic cleft are known as the ‘extra-synaptic’

21 receptors. Receptors containing a γ2-subunit in association with α1-, α2- or α3- subunits (α1β2/3γ2, α2β2/3γ2 and α3β2/3γ2) are the predominant synaptic receptor subtypes (Farrant and Nusser, 2005). Receptors that contain α4-, α5- or α6- subunits (α6βnδ, α4βnδ and α5βnγ2) are predominantly or exclusively extrasynaptic (Fritschy and Brunig, 2003; Nusser et al., 1998). In synapse, each vesicle is thought to liberate several thousand GABA molecules into the synaptic cleft, produce a response (IPSC) at high concentration of GABA (0.3-1.0 mM) and they are short lived (10-100 ms) (Mozrzymas et al., 2003; Semyanov et al., 2004). There are only few number of receptors (from ten to a few hundred) clustered opposite the release site (Brickley et al., 1999; Mody et al., 1994; Nusser et al., 1997). Synaptically released GABA acting on postsynaptic GABAA receptors produces “phasic” inhibition, whereas “tonic” inhibition results from the continuous activation of extrasynaptic receptors by ambient GABA (Farrant and Nusser, 2005). The main feature of phasic GABAA receptor-mediated inhibition is the rapid synchronous opening of a relatively small number of channels that are clustered at the synaptic junction, whereas tonic inhibition results from random, temporally dispersed activation of receptors that are distributed (albeit in a potentially non-uniform manner) over the neuronal surface. GABA-mediated tonic conductance are found in granule cells of the dentate gyrus (Farrant and Nusser, 2005; Nusser and Mody, 2002), thalamocortical relay neurons of the ventral basal complex (Porcello et al., 2003), layer V pyramidal neurons in the somatosensory cortex (Yamada et al., 2004), CA1 pyramidal cells (Bai et al., 2001), and certain inhibitory interneurons in the CA1 region of the hippocampus (Semyanov et al., 2003). Unlike receptors mediating phasic current, tonically active GABAA receptors show unusual high GABA affinity (Saxena and Macdonald, 1996) and be activated by the low ambient GABA concentrations (nanomolar to few micromolar) (Lerma et al., 1986; Tossman et al., 1986). At submicromolar concentrations, several competitive and non-competitive GABAA antagonists (, picrotoxin and bicuculline) reduced phasic currents, but had no effect on tonic currents. However, the antagonists blocked both phasic and tonic currents at high concentrations (Bai et al., 2001; Farrant and Nusser, 2005; Semyanov et al., 2003; Stell and Mody, 2002; Yeung et al., 2003). There is also difference in the tonic and phasic inhibitions by neuroactive steroids which will be discussed later. GABA receptor-mediated inhibitory postsynaptic currents (IPSCs) peak rapidly (0.5-5 ms) and usually decay with two time constants (slow and rapid decay) of ranging from few millisecond to tens or hundreds of milliseconds (Jones and Westbrook, 1995; Mozrzymas et al., 2003). A commonly observed use-dependent characteristic of GABAA receptors is desensitization. Desensitization can be defined macroscopically; as the 'decline in response in the continued presence of agonist'. Microscopically, it is essentially a 'long-lived agonist-bound closed state

22 of the receptor'. As ‘desensitized’ states are part of the generally agreed reaction mechanism for the GABAA receptor, exposure of the receptor to GABA will drive entry to these states; the time spent in these states will reflect the waveform of GABA exposure. With voltage-clamp of oocytes expressing GABAA receptor, the ability to apply agonist quickly is severely limited by the physical size of the oocyte, and so desensitization happens during the slow onset phase. This means that information relevant to the rapid onset of the IPSC is lost. Although such experimental stimuli are very different from the natural stimulus experienced by GABAA receptors in the brain, several physiologically relevant aspects of inhibition can be derived from this fundamental receptor property. As mentioned earlier the time course of the GABA concentration is transient and the decay of the IPSC is dominated by the ion channel closure that follows ligand removal, a macroscopic phenomenon known as deactivation.

4.8 Pharmacological effects depend on GABA receptor subunit composition There are evidences that several subunits are related to several behavioral effects. Transgenic mouse models enable exciting new perspectives for the development of pharmacology of modulating agents of GABAA receptor and their effect on different subunits. Benzodiazepines have two effects in the GABA system- sedative and anxiolytic. Histidine residue in α1, α2, α3, and α5 subunits is crucial for benzodiazepines binding. Sedative effect of benzodiazepines is mediated via the α1 subunit containing GABAA receptors (Rudolph et al., 1999). Replacement of histidine by arginine at position 101 at α1 subunit leads to mice in which the anxiolytic and muscle relaxant effects of diazepam are retained, whereas its effects on the motor component of sedation, its amnesic and anticonvulsant effects are abolished (McKernan et al., 2000; Rudolph et al., 1999). In contrast, the α2 subunit appears to be a major determinant for the anxiolytic and muscle relaxant effects of benzodiazepines (Crestani et al., 2001; Low et al., 2000). Agonistic activity at the α3 subunit also mediate the anxiolytic activity of benzodiazepines, although it occurred only at high receptor occupancy (Dias et al., 2005). Recently, it has been shown that the α5 subunit is important for sedative tolerance development to benzodiazepines and for acquisition and expression of associative memory and spatial learning (Collinson et al., 2002; Crestani et al., 2002; van Rijnsoever et al., 2004; Yee et al., 2004). In addition, α4 subunit is also implicated in the regulation of (Gulinello et al., 2001). A concentration-dependent decrease of the α4 subunit is seen after 4-day application of allopregnanolone to developing neuronal cells (Grobin and Morrow, 2000), whereas in the hippocampus and cerebellum, an increase in this subunit can be detected after withdrawal from chronic progesterone (or allopregnanolone) exposure and after short-term treatment (Concas et al., 1999; Follesa et al., 2001; Gulinello et al., 2001; Smith et al., 1998). α6 subunit is highly sensitive to pentobarbital effect

23 (Thompson et al., 1996) and neuroactive steroid (Belelli et al., 2002). In recent years, there has been an accumulating amount of evidence for the involvement of GABAA receptors in the action of anesthetics. The intravenous anesthetic etomidate shows GABAA subtype selectivity in vitro for β2- and β3-containing receptors when studied on recombinant receptors (Hill-Venning et al., 1997). A recent study shows that β3-containing receptors are the primary mediators of the anesthetic effects whereas β2, probably in combination with α1, and γ subunits mediates the sedating effects of etomidate (Wafford et al., 2004). The γ-subunit is thought to confer sensitivity of the receptor to benzodiazepines (Pritchett et al., 1989; Ymer et al., 1990). The γ1- and γ3-containing receptors have a 10-30 fold lower affinity for than do receptors containing γ2-subunit (Hadingham et al., 1995; Luddens et al., 1990; Wafford et al., 1993; Ymer et al., 1990). On the other hand, absence of γ subunit is essential for Zinc sensitivity (Hosie et al., 2003). The γ2-subunit is also involved in anxiety regulation, and it is changed during hormone manipulation and pregnancy (Concas et al., 1999; Crestani et al., 1999; Essrich et al., 1998; Follesa et al., 1998; Kittler et al., 2000). The γ2-subunit is essential for clustering of GABAA receptors and gephyrin and synaptic localization (Baer et al., 1999; Essrich et al., 1998) (Fritschy and Brunig, 2003; Schweizer et al., 2003). The δ-subunit is responsible for tonic conductance and important for neuroactive steroid modulation on GABAA receptor (Stell et al., 2003). Receptor knockout studies have revealed that the absence of the δ-subunit decreases the sensitivity to neuroactive steroids such as pregnanolone and alphaxalone, thereby influencing the duration of anesthesia and the anxiolytic effect of those steroids (Mihalek et al., 1999). The ε-subunit reduces neuroactive steroid and anesthetic modulation (Belelli et al., 2002; Davies et al., 1997; Thompson et al., 2002). Functionally, distinct subunit-specific properties have been identified in both recombinant and native receptors, supporting the concept that GABAA receptor heterogeneity is a major facet determining the functional properties of GABAergic inhibitory circuits (Mohler et al., 2001; Sieghart, 2000). In particular, the type of α- subunit determines the kinetics of receptor deactivation (Devor et al., 2001; Hutcheon et al., 2000; Verdoorn et al., 1990), and the presence of the δ-subunit results in markedly increased agonist affinity and apparent lack of desensitization (Adkins et al., 2001; Burgard et al., 1996; Fisher and Macdonald, 1997).

4.9 GABAA receptor and disorders When the balance between excitatory and inhibitory activity is shifted pharmacologically towards GABAergic transmission anxiolysis, sedation, amnesia, ataxia, and loss of consciousness can be induced. On the other hand, an attenuation of the GABAergic system results in arousal, anxiety, restlessness, , exaggerated reactivity, and even epileptic seizures. These GABAergic

24 system is known to multiple neurological and psychiatric diseases, including anxiety disorders (Malizia, 1999), epilepsy (Backstrom et al., 2003; Coulter, 2001; Duncan, 1999; Olsen et al., 1999), ethanol dependence (Morrow et al., 2001), Huntinton’s disease (Kunig et al., 2000), Angelman syndrome (DeLorey et al., 1998), and schizophrenia (Blum and Mann, 2002; Lewis, 2000; Nutt and Malizia, 2001). There is strong evidence that neuroactive steroid allopregnanolone and 3α5αTHDOC are involved in the pathophysiology of premenstrual syndrome, catamenial epilepsy, major depression, and stress-sensitive brain disorders. Neuroactive steroids PS and DHEAS have been shown to modulate memory functions. However, the significance of the testosterone-derived neuroactive steroid 3α-androstanediol is not well understood (Reddy, 2003).

5. Neuroactive steroid agonists Recent studies have indicates the existence of at least three effects of neuroactive steroids on GABAA receptor: one for the allosteric enhancement of GABA- evoked current, one for direct activation and one for the antagonistic action by the sulfated and 3β-OH steroids.

5.1 Synaptic effect of agonist neuroactive steroids As stated earlier GABA mediated inhibition may be of two varieties; synaptic and extrasynaptic. The brief inhibitory response of steroids that results from the activation of postsynaptic GABAA receptor is phasic response. Although both modes of operation clearly impact upon neuronal information processing at the cellular and network level, the extent to which each receptor pool influences brain excitability in normal and diseased states is not known. Synaptic GABAA receptors are ternary complexes that almost invariably incorporate the γ2-subunit in combination with α (mainly α1-, α2- and α3-) and β2/3-subunit isoforms. However, these receptor isoforms can also be located extra-synaptically (Farrant and Nusser, 2005). The kinetic of agonist steroids at synaptic GABAA receptor has been studied thoroughly by measuring the spontaneous inhibitory postsynaptic currents (sIPSC) from neurons in brain slices. The neurosteoids have little effect on the rise time or peak amplitude of the sIPSC, but they prolong the decay of IPSC (Haage et al., 2005). However this effect is neuron specific. In hippocampal CA1 neurons, cerebellar granule cells and Purkinje neurons, neuroactive steroids prolong the sIPSC at relatively low concentration (in nanomolar range) (Cooper et al., 1999; Harney et al., 2003; Vicini et al., 2002). On the other hand, micromolar concentrations are required to produce equivalent responses in hypothalamic neuron (Brussaard et al., 1997; Koksma et al., 2003). Alphaxalone enhance GABAA receptor modestly in hippocampus and not at all in the caudate nucleus. It enhances GABA binding in the temporal lobe better than the parietal and frontal lobe with no effect in the occipital lobe (Nguyen et al., 1995).

25 3α5αTHDOC enhanced GABA binding better in CA3 and subiculam than CA1 and entorhinal cortex (Nguyen et al., 1995).

Figure 9. Neuroactive steroids’ effect on the synaptic and extrasynaptic GABAA receptors and type of inhibition produced by them (adapted from Belelli and Lambert, 2005).

The effect of neuroactive steroids also depend on concentration (Haage and Johansson, 1999). In the dissociated neurons from the medial preoptic nucleus, 2µM 3α5αP modulating 100 µM GABA response is markedly depressed and the desensitization is faster, but the decay after GABA application is prolonged. In contrast, modulating 1 µM GABA response is markedly potentiated, the activation is faster, a prominent desensitization is induced, and the decay after GABA application is prolonged. Neurons in the same brain region can also show the heterogeneity. In hippocampal brain slices from 20-day-old rats, an approximately 30-fold higher concentration of pregnanolone (3α5βP) is required to prolong mIPSCs in dentate granule cells (DGCs) than in CA1 (Belelli and Lambert, 2005; Harney et al., 2003). Neuroactive steroid sensitivity to GABAA receptor of different brain regions also depend on the different stages of development. In

26 DGCs, mIPSCs that are recorded from 10-day-old rats are more sensitive than those from 20-day-old animals (Cooper et al., 1999). Recorded from HEK293 cells transfected with synaptic receptors and recombinant synaptic GABAA receptors showed that 3α5αP, 3α5αTHDOC and ACN ((3α5α17β)-3-hydroxyandrostane-17-carbonitrile) and B285 (3α5α17β)-3- hydroxy-18-norandrostane-17-carbonitrile) are the positive modulators of GABA site (Akk et al., 2004; Belelli et al., 2002; Reddy, 2003). These potentiating steroids prolong GABAergic inhibitory postsynaptic currents, although the amplitude is generally not increased (Haage et al., 2005). 3α5αP and probably other potentiating steroids are assumed to reduce the rate of GABA unbinding from the receptor (Haage et al., 2005).

At higher concentration (>10 µM), which occur during parturition (Stoffel- Wagner, 2003), neuroactive steroids directly activate the receptor (Majewska et al., 1986) similar to those of barbiturates, steroids but interact with different sites on GABAA receptor (Kerr and Ong, 1992). This ‘GABA-mimetic’ effect of neuroactive steroids is sufficient to suppress the excitatory neurotransmission (Shu et al., 2004).

5.2 Extrasynaptic effect of agonist neuroactive steroids The response of neuroactive steroids at relatively low concentration is mediated by the activation of extrasynaptic or perisynaptic α4, α5, α6 and δ-subunit containing GABAA receptors (Fig 9). These extra-synaptic receptors identified in hippocampal dentate gyrus granule cells (DGCs), the granule cells of the cerebellum and the relay neurons of the thalamus are distinct from their synaptic counterparts. Extrasynaptic conductance can have a considerable influence on neuronal excitability (Leroy et al., 2004). Such receptor assemblies exhibit both a high affinity for GABA and a reduced receptor desensitization in the continued presence of the agonist, properties that render these receptors ideally suited to sense the low ambient concentrations (~0.5-1 μM) of the estimated to be present extra-synaptically (Kennedy et al., 2002). These receptors are highly sensitive to neuroactive steroids in certain brain region. Low ‘physiological’ concentrations (10-100 nM) of 3α,5α-THDOC in mouse DGCs and CGCs, selectively enhance the tonic conductance, with little or no effect on the phasic conductance (Belelli and Lambert, 2005; Farrant and Nusser, 2005; Stell et al., 2003). GABAA receptor δ-subunit in a ternary combination (αβδ) contains α6- and α4-subunits (Farrant and Nusser, 2005; Fritschy and Brunig, 2003; Stell et al., 2003). Recombinant receptors containing the δ-subunit are highly sensitive to neuroactive steroid modulation (Wohlfarth et al., 2002). Tonic inhibition is reduced in δ subunit ‘knock-out’ mice, and the residual tonic current was insensitive to 3α5α-THDOC (Mihalek et al., 1999; Stell et al., 2003). The inhibitory

27 effect of neuroactive steroids differs in different brain region. 250 nM 3α5αTHDOC has no effect on the tonic inhibition in ventrobasalis complex of the thalamus (Porcello et al., 2003). Similarly, the tonic conductance of hippocampal CA1 neurons, which is mediated partly by receptors that contain the α5-subunit, is affected by 3α5αTHDOC above 100 nM (Belelli and Lambert, 2005; Farrant and Nusser, 2005; Stell et al., 2003). The modulation of tonic GABAA- receptor-mediated currents by neuroactive steroids is influenced by local neuroactive steroid metabolism, inhibition of which can greatly enhance the response of the tonic current in dentate granule cells to endogenous neuroactive steroids, while having no effect on the response to the synthetic metabolically stable neuroactive steroid ganaxalone (Belelli and Herd, 2003). Such neuroactive steroid metabolism demonstrates regional specificity, which could further contribute to the regional specificity of the tonic GABAA-receptor-mediated currents, and their modulation. This raises the intriguing possibility of selectively regulating network excitability with inhibitors of neuroactive steroid metabolism; these could thus have an antiepileptic action. In summary, evidence is emerging that the GABA mediated tonic conductance present in some neurons may have a considerable influence on neuronal signaling and network activity (Brickley et al., 2001; Hamann et al., 2002; Mitchell and Silver, 2003). The high sensitivity of neuroactive steroids in extrasynaptic receptor may represent an important target for these steroids.

Using labeled α- (Bgt) derivatives, it was observed in hippocampal neurons that the principal sites of both GABAA receptor endo- and exocytosis are extrasynaptic (Bogdanov et al., 2006). In addition, the synaptic GABAA receptors are recruited directly from their extrasynaptic counterparts, and constitute a dynamic mechanism for neurons to rapidly modulate receptor number at inhibitory synapses by controlling the availability and stability of extrasynaptic receptors.

5.3 Subunit dependence of neuroactive steroids As mentioned earlier that the GABA-ergic synaptic transmission may be differentially regulated by pregnane steroids in different brain regions, an effect that might be attributable to variations in GABAA receptor subunit composition. The α-subunit did not greatly influence the GABA-modulatory actions of 3α5αP, when co-expressed with β1 and γ2L subunits in Xenopus oocytes (Belelli et al., 2006). Likewise, the isoform of the β-subunit (1-3) has little influence on the GABA-modulatory actions of the pregnane steroids (Belelli et al., 2002; Hadingham et al., 1993; Sanna et al., 1997). The presence of γ-subunit is not a prerequisite for neuroactive steroid activity. In fact, the efficacy of 3α5αP modulation mediated by binary α1β1 is higher than that mediated by ternary

28 α1β1γ2 receptors. That is, the steroid to be more effective at the binary isoform, increasing the GABA-evoked response above the apparent maximal response to GABA (Belelli et al., 2002; Maitra and Reynolds, 1999). The isoform of the γ- subunit has little, or no, effect on the maximal GABA-modulatory effect of 3α5αP but significantly influences the potency of the steroid with “physiological concentrations” (3-30 nM). On the other hand, δ-subunit when coexpressed with α4- and β3- subunits, a receptor thought to be naturally present in the thalamus (Sur et al., 1999) shows high steroid sensitivity compare to γ- subunit containing receptor (Belelli et al., 2002; Davies et al., 1997). Receptors incorporating the ε- subunit are reported to be insensitive to the modulatory actions of the pregnane steroids, not the direct GABA-mimetic effect.

5.4 Structure activity relationship Structure-activity relationship of the neuroactive steroids and the effects associated with their use has been summarized in a number of articles (Hamilton 2002; Gyermek et al., 1968; Gyermek and Soyka, 1975; Laubach et al., 1955). The systemic investigation of several isomer combination of GABAA-active neuroactive steroids revealed, as a general rule that several features are of importance for an effective drug: geometry between ring A/B is crucial, Hydrogen bond donator in position 3, hydrogen-bond acceptor in position 20 and/or flexible bond at position 17 (Ragagnin et al., 2007, Purdy et al., 1990; Zorumski et al., 2000). By convention, the α- and β-refer to substituents below and above the plane of the steroid rings, respectively. In the next paragraphs, we describe the variation in the core and substitution in certain positions on the steroid nucleus. Modification of C3 position is described in the neuroactive steroid antagonists section.

5.4.1 Backbone structure of neuroactive steroids Anesthetic steroids typically have a saturated backbone of four rings, although this is not an absolute requirement for activity. These rings form a rigid framework for positioning the hydrogen bonding groups in three-dimensional spaces. Certain benz[e]indines have steroid like effect on neuron (Hamilton, 2002; Rodgers- Neame et al., 1992), even though they are tricyclic steroid analogues in which the steroid-A ring is opened and partially removed. The presence of hydrogen bond donor in the α-configuration at C3 and β-configuration at C17 are critical for anesthetic actions (Hawkinson et al., 1994; Hogenkamp et al., 1997; Purdy et al., 1990). These groups are important for the binding of steroids to a variety of proteins by means of hydrogen-binding with polar or charged residues (Brzozowski et al., 1997; Grishkovskaya et al., 2000). Replacing the hydrogen of hydroxyl with methyl, thus eliminating the ability of the steroid to donate a

29 hydrogen bond in this region, results in much reduction in potency (Upasani et al., 1997).

5.4.2 Steroid- Mimetic Core Replacing the steroid skeleton with an alicyclic framework gave a compound that showed weak potentiating activity (Burden et al., 1998). When the nitrile is replaced with an acetyl group the derivative became inactive (Hamilton, 2002).

5.4.3 Replacing C5-position The configuration at C5 is also important. Even if steroids with either 5α or 5β conformations are active, spatial difference in this position may affect the pharmacology of the neuroactive steroids. Studies with 5α-THDOC and its stereoisomer 5β-THDOC revealed important differences in potency, efficacy and regional selectivity at the GABAA receptor complex (Gee and Lan, 1991; Mennerick et al., 2004). Moreover, 5α reduced steroid, but not 5β-steroids, show a high degree of enantioselectivity/enantiospecificy in their action as modulator of GABAA receptor and as anesthetics (Covey et al., 2000). Although model has been proposed to account for steroids binding at the same site by invoking conformational changes (Hamilton, 2002; Purdy et al., 1990), it is also possible that they occupy different sites favored by their respective stereochemistries. Removal of angular methyl group of 5α -steroids i.e. 19-nor derivatives has variable effect on the potentiating activity of 5α and 5β-steroids, whereas incorporation of a methyl group at C5 eliminate or diminishes activity (Han et al., 1996).

5.4.4 Replacing C11-position The very potent but opposing effects of picomolar concentrations of cortisol and on GABA receptors in the intestine, with cortisol enhancing and cortisone reducing GABA responses, is very interesting both from a structure- activity and a physiological viewpoint. Cortisol and cortisone differ in structure only by the level of oxidation at carbon 11, cortisol being the 11β-hydroxy compound and cortisone the 11-oxo compound. 11-ketones like alphaxalone and 11α-amines i.e. minaxalone retain good activity and this is one of the few cases where polar groups are tolerated on the α-face of the steroid without serious loss potentiating activity (Hamilton, 2002). However in chloride uptake studies in cortical synaptosomes the 3α5α metabolite of cortisol and cortisone are antagonistic but enhance the positive allosteric effect of allopregnanolone (Stromberg et al., 2005).

30 5.4.5 Replacing C20-position Pregnan-20-one is easily metabolized to pregnan-20-ols. Reduction to 20α-or 20β- ol reduces the affinity and efficacy of pregnan steroids. However, a ketone is preferred 20-substituent for pregnanes for good its efficacy (Belelli et al., 1996; Hamilton, 2002; McCauley et al., 1995).

5.4.6 Replacing C21-position A variety of substituents can be accommodated at the C21-position, i.e. halides (Hill-Venning et al., 1996), (Xue et al., 1997) and thiosulfate (Fick et al., 1999) and heterocycles (Yang et al., 1999), generally without serious loss of activity. 3α5αTHDOC is a potent and efficacious GABAA , whereas the 21-OH analogues of pregnenolone are not. It rather antagonizes the effect of 3α5αP (Hamilton, 2002 ; Xue et al., 1997).

5.4.7 Steroid nucleus Andros-5-enes and pregn-5-enes with polar functional groups at the 17β- and C20 positions respectively retain good inhibitory activity (Hamilton, 2002).

5.5 Binding sites of neuroactive steroids Determining how neuroactive steroids interact with the GABAA receptor is a prerequisite for understanding their physiological and pathophysiological roles in the brain. Neuroactive steroids bind to GABAA receptors at a site that is distinct from the recognition sites for GABA, benzodiazepines, and barbiturates. This results in allosteric modulation of GABA binding or channel gating. Electrophysiological studies have confirmed that neuroactive steroids enhance chloride current through increasing both channel frequency and channel open duration (Callachan et al., 1987; Puia et al., 1990; Zhu and Vicini, 1997). Neuroactive steroids do not require direct aqueous access to the receptor, and membrane accumulation is required for receptor modulation (Akk et al., 2005). In a recent study, Hosie et al. (2006) identified two discrete binding sites in the receptor's transmembrane domains that mediate the potentiating and direct activation effects of neuroactive steroids. Their potentiating effect is mediated by a cavity formed by the α-subunit transmembrane domains. On the other hand, their direct activation is mediated by interfacial residues between α and β-subunits and is enhanced by steroid binding to the potentiation site. These profiles indicate that two distinct neuroactive steroid binding sites may exist; αTHr236 and βTyr284 residues in the transmembrane domain initiate direct activation whereas αGln241 and αAsn407 mediate the potentiating response (Hosie et al., 2006).

31 5.6 Physiological role of neuroactive steroids Many of the physiological actions of neuroactive steroids at the GABAA receptor complex have been evaluated (Majewska, 1987). As the neuroactive steroids are converted from gonadal hormone, it is often difficult to establish if the effect is due to parents hormonal steroids or their metabolites (Rupprecht et al., 1996). However, virtually all enzymes mediating these transformations have been identified in the CNS (Mellon et al., 2001). The concentrations of circulating neuroactive steroids vary during menstrual cycle, pregnancy (Bixo et al., 1997; Herbison, 2001), delivery and stress (Purdy et al., 1991). Therefore there is an obvious potential links with mood disorders such as stress induced depression and cognitive disturbance, PMDD and post-natal depression. In addition to modulating receptor function, there is evidence that treatment with steroids induce receptor plasticity through changes in expression of particular GABAA receptor subunits (Follesa et al., 2001; Gulinello et al., 2001). Following prolong administration pregnenolone causes expression of neuropeptide Y receptor Y1 gene in the amygdala, a region in the brain that is important for controlling anxiety (Ferrara et al., 2001). This is similar to the effect seen with ligands that bind to the benzodiazepine site of GABAA receptor and supports the hypothesis that there is a functional interaction between GABA and neuropeptide Y in the amygdala. Moreover, the endogenous benzodiazepine octadecaneuropeptide regulates corticotrophin-releasing hormone (CRH) mRNA expression, and this regulation is mediated by the effects of the peptide at GABAA receptors and positively influenced by the neuroactive steroids (Givalois et al., 1998). There is also evidence that allopregnanolone, at nanomolar concentration in vitro suppress the release of hypothalamic gonadotropine releasing hormone (Calogero et al., 1998). Other physiological actions mediated by the neuroactive steroids include the prevention of protein kinase C-dependant GABAA receptor in oxytocine neurons at the juvenile stage or during late pregnancy; after parturition GABAA receptors in oxytocine neurons are less sensitive to allopregnanolone (Brussaard et al., 2000); and possible effects on neurotransmission due to the catalepsy observed in mice following administration of allopregnanolone (Khisti et al., 1998). Side effects typical of benzodiazepines, such as tolerance, dependence and interactions with alcohol are also observed in prolong treatment of neuroactive steroids (Morrow et al., 2001). However, synthetic analogues appear to offer some advantages over the endogenous steroids with respect to their potential for abuse (Rowlett et al., 1999). Neuroactive steroids effect has also been reported in (NMDA) and sigma receptors (Baulieu, 1998; Mellon et al., 2001; Reddy and Kulkarni, 2000; Zorumski et al., 2000). Other actions are liable to be non-GABAergic include the suppression of pituitary follicle stimulating hormone from rat pituitary cells by 3α-hydroxy-4-pregnen-20-one and the blockage of voltage activated calcium channels in rat hippocampal and dorsal root

32 ganglion neurons by 3α-hydroxy-5α-androstane-17β-carbonitrile (Nakashima et al., 1998).

5.7 Neuroactive steroids as potential drugs A number of review articles have covered the potential that neuroactive steroids modulating GABAA receptor system might offer for treating various disorders (Majewska, 1992; Reddy and Kulkarni, 2000; Rupprecht et al., 2001; Zorumski et al., 2000). However, certain obstacles prevent the clinical use of endogenously occurring neuroactive steroids. Importantly, natural neuroactive steroids such as allopregnanolone have low bioavailability because they are rapidly inactivated and eliminated by glucoronide or sulfate conjugation at the 3α-hydroxy group. In addition, the 3α-hydroxy group of allopregnanolone may undergo oxidation to the ketone, restoring activity at receptors (Rupprecht et al., 1993). Ganaxalone (3α-hydroxy-3β-methyl-5α-pregnane-20-one), the 3β-methyl analogue of allopregnanolone, is an example of a synthetic neuroactive steroid congener that overcomes these limitation (Carter et al., 1997). Like allopregnanolone, ganaxalone is a positive of GABAA receptor. The following section will therefore focus on the therapeutic potential of the neuroactive steroids.

5.7.1 Anxiolytics effect of neuroactive steroids Neuroactive steroids modulate relatively to anxiety and stress. It has been proposed that their anxiolytic activity is a consequence of regulation of steroid biosynthesis via the hormones adrenocorticotropin hormone (ACTH) and corticotropine releasing hormones (CRH) (Torres et al., 2001). After an acute stress stimulus there is also a release of progesterone, pregnenolone, allopregnanolone and 3α5β-THDOC (Barbaccia et al., 1998; Purdy et al., 1992; Serra et al., 2000). Allopregnanolone and 3α5αTHDOC have been demonstrated to possess potent anxiolytic activity in several different animal anxiety models (Bitran et al., 1995; Crawley et al., 1986; Reddy and Kulkarni, 2000; Wieland et al., 1995). The anxiolytic effect has however never been shown in humans (Wihlback et al., 2006). A number of recent reports have actually indicated that allopregnanolone can induce aggression and anxiety (Fish et al., 2001; Gulinello et al., 2001; Miczek et al., 2003). A recent clinical study showed that the allopregnanolone and cortisol level are increased during the examination of PhD students (Droogleever Fortuyn et al., 2004). Therefore it has been suggested that allopregnanolone in certain individuals has biphasic effects; with low doses increasing an adverse, anxiogenic effect, and high doses decrease this effect and show beneficial, calming property. A major concern with potential new anxiolytics is whether they suffer the same drawbacks as classical benzodiazepine. Using the elevated plus-maze paradigm for assessing the anxiolytic activity, behaviorally

33 selective effects of neuroactive steroids have been reported which differ from those of the benzodiazepine diazepam (Rodgers and Johnson, 1998). Evaluation of novel neuroactive steroids as anxiolytics continues. Synthetic derivative Co 2- 6749 30, retains a 3β-trifluromethyl group that should block metabolism and enhance oral bioavailability, was selected for clinical development because there is a large separation between anxiolytic-like effect and side effects (Gasior et al., 1999; Vanover et al., 2000).

5.7.2 Effect of neuroactive steroids in learning and memory Hippocampus is a key brain area for learning and memory functions (Farr et al., 2000). Women with premenstrual dysphoric disorder often show difficulties in concentrating and develop fatigue during the luteal phase of the menstrual cycle; this is associated with high circulating levels of allopregnanolone (Sundstro and Backstrom, 1999). The enzymes needed for the production of allopregnanolone are also present in certain areas of the brain, notably the hippocampus (Compagnone and Mellon, 2000). Using the Morris water maze paradigm, allopregnanolone was found to inhibit learning (Johansson et al., 2002). 3β-steroid namely, 3β-20β-dihydroxy-5α-pregnane, antagonism at the GABAA receptor reduces the negative allopregnanolone effect on learning in the water maze (Turkmen et al., 2004). Pregnenolone sulfate infused into the basal magnocellualar nucleus enhance memory performance, whereas allopregnanolone disrupted memory (Mayo et al., 1993). Pregnenolone, DHEA and DHEAS increased memory when injected systemically, centrally or into the amygdala (Flood et al., 1992; Flood et al., 1988; Wolkowitz et al., 1995). There is evidence that the concentration of DHEA and DHEAS are decreased in patients suffering from Alzheimer’s disease (Hillen et al., 2000; Nasman et al., 1991; Sunderland et al., 1989). Thus neuroactive steroids should be further explored in the context of prevention of and/or treatment of age-related memory disorders.

5.7.3 Effect of neuroactive steroids on sleep As early as 1942, Selye reported the sedative and anesthetic properties of progesterone and some of its metabolites (Selye, 1942). 3α5αP, 3α5βP, 3α5αTHDOC exhibit benzodiazepine like effect, including reduced sleep latency and increased non-REM sleep with only small changes in slow wave and REM sleep (Bowers and Wehner, 1992; Olney et al., 1991). The limited bioavailability of steroids following oral administration is a significant issue in the development of these compounds as drug. However, several new compounds have been evaluated in animal models as potential and ; side effects for these activities are liable to be less severe than for anesthetics due to the reduced doses and a wider therapeutic margin. Tolerance to minaxadole and allopregnanolone has been reported at sedative doses following chronic treatment (Marshall et al.,

34 1997; Turkmen et al., 2006). Other compounds that have been evaluated as sedative hypnotics include Co 134444, Co 177843 and 177834 and Co 127501(Hamilton, 2002; Vanover et al., 1999; Vanover et al., 2001), CCD-3693 31 (Edgar et al., 1997), CCD-3693 is undergoing clinical evaluation for insomnia (Gasior et al., 1999).

5.7.4 Effect as anticonvulsant Neuroactive steroids are potent broad-spectrum anti-seizure agents. Several neuroactive steroids protect against seizure induced in animal GABAA receptor antagonist, picrotoxin induced seizures, and kindled seizures (Reddy, 2003; Reddy et al., 2004; Turner et al., 1989; Wieland et al., 1995). Some neuroactive steroids are highly effective in suppressing -,ethanol-, diazepam-, and neuroactive steroid-withdrawal seizures (Devaud et al., 1996; Devaud et al., 1995; Reddy et al., 2001). At very high doses, the neuroactive steroids also partially protect mice against maximal electroshock-induced seizures. Similarly, synthetic allopregnanolone analogues demonstrate comparable anticonvulsant efficacy (Carter et al., 1997; Hogenkamp et al., 1997). Ganaxalone has activity in a broad range of animal models of epilepsy. It has been shown to be well tolerated in adults and children (Nohria and Giller, 2007). The clinical studies with ganaxalone suggest the dose limiting side effect is sedation (Beyenburg et al., 2001). Withdrawal (Reddy and Kulkarni, 2000; Reilly et al., 2000), tolerance (Czlonkowska et al., 2001) and sedation (Kokate et al., 1994) are the associated with and neuroactive steroids that have been observed in rodent seizure model.

5.7.5 Analgesic effect of neuroactive steroids Recent reports indicate alphadalone and water soluble-amino steroids and their metabolite to be analgesic (Hamilton, 2002). However, related compound alphaxalone, which differs only in lacking a 21-OH group, is ineffective (Goodchild et al., 2001; Nadeson and Goodchild, 2001). The analgesic activity has been attributed to GABAA receptors in the spinal cord since the antinociception is reversed by administration of the GABAA antagonist bicuculline (Hamilton, 2002).

6. Neuroactive steroid antagonists 6.1 Pregnenolone sulfate other sulfated steroids 6.1.1 Modulation of GABAA receptor by pregnenolone sulfate Sulfated steroids like pregnenolone sulfate (PS), and dehydroepiandrosterone sulfate (DHEAS) can produce profound effects on behavior. PS, an abundant neuroactive steroid, enhances learning (Flood et al., 1995; Mayo et al., 1993) while antagonizing the impairment of learning and memory produced by ethanol and

35 (Melchior and Ritzmann, 1996). PS may play a role in cognition and have been reported as negative modulators of GABA receptors based on electrophysiological studies and GABA mediated 36Cl- uptake by rat brain synaptosomes (Demirgoren et al., 1991; Majewska et al., 1990). Although structural similarity of pregnenolone with that of 3α5α-steroids might lead one to expect that PS would share a common binding site with 3α5αP, this is not the case. Structure activity relationships for GABAA receptor modulation are different for sulfated inhibitory steroids . nonsulfated potentiating steroids (Park-Chung et al., 1999). Potentiation by nonsulfated steroids requires 3α stereochemistry. Pregnenolone is inactive at GABAA receptors. In contrast, PS is inhibitory, both as the 3α and 3β isomers. Although the addition of a negatively charged sulfate or hemisuccinate group at the C3 position converts a number of neuroactive steroids from potentiating to inhibitory (Park-Chung et al., 1999), a negatively charged group at C3 is not absolutely essential for inhibition, as the nonsulfated neuroactive steroid dehydroepiandrosterone (DHEA) is also inhibitory, although less potent than its sulfated derivative DHEAS (Imamura and Prasad, 1998; Park- Chung et al., 1999). Some steroids, such as 11-keto derivative of PS is of particular interest, as it can behave as a positive, negative, or neutral modulator. It was suggested that 11-ketopregnenolone sulfate (11-ketoPS) exerts a dual action on distinct positive and negative steroid modulatory sites associated with the GABA receptor (Park-Chung et al., 1999).

6.1.2 Mechanism of action of pregnenolone sulfate The mechanism of inhibition by pregnenolone sulfate has been studied extensively. Inhibition of GABAA receptor function by PS and DHEAS is not voltage-dependent (Gibbs et al., 2006; Majewska et al., 1988; Spivak, 1994), indicating that these sulfated steroids do not need to penetrate significantly into the membrane field to reach their sites of action, and suggesting that they act allosterically rather than by physically occluding the ion pore of the GABAA receptor. Pregnenolone sulfate has been reported to act by reducing channel- opening frequency (Mienville and Vicini, 1989). A subsequent single-channel patch clamp study of α1β2γ2L GABAA receptors expressed in HEK293 cells found that PS had no effect on the rate constants for channel closing and opening, and concluded PS produces a block that is independent of GABA binding. Channel activation state, or membrane potential, with an IC50 of 100 nM, suggesting that PS acts essentially as a slow noncompetitive antagonist (Akk et al., 2001). In contrast, studies of macroscopic GABAA receptor currents have reported that PS modulates synaptic transmission by enhancing GABA receptor desensitization (Haage et al., 2005; Shen et al., 2000). The disagreement between microscopic and macroscopic studies of PS inhibition of GABAA receptor function remains to be resolved (Gibbs et al., 2006).

36

6.1.3 Site of action of pregnenolone sulfate The site of action of sulfated neuroactive steroids on GABAA receptors remains unclear. Based upon the observation that PS reduced the apparent affinity of 35S- TBPS, Sousa and Ticku (1997) suggested that PS and DHEAS might bind at the picrotoxin/cage convulsant site. However, a mutation to the transmembrane M2 channel domain eliminated picrotoxin sensitivity but the inhibitory effects of PS and DHEAS persisted (Gibbs et al., 2006; Shen et al., 1999). The absence of voltage sensitivity or alteration of single-channel open time argues against a binding site within the pore. Akk et al. (2001) identified a valine residue in the channel domain of the α1 subunit that slowed the development of PS inhibition when mutated to , but concluded that this residue is unlikely to be part of the binding site and likely influences PS action indirectly (Akk et al., 2001). A recent study in C. elegans identified multiple residues in transmembrane domain 1 (M1), as well as a residue near the extracellular end of the M2 helix, that are critical for low-μM inhibition of C. elegans GABAA receptors by PS. This latter residue is of particular interest, as it is a positively charged arginine that could potentially coordinate with the negatively charged sulfate of PS. The C. elegans receptor exhibits some pharmacological differences as compared to mammalian GABAA receptors (for example, pregnanolone is inhibitory), so these results may or may not be relevant to mammalian receptors; however, it is notable that an arginine residue is also found in this region of mammalian GABAA receptor subunits (Gibbs et al., 2006; Wardell et al., 2006).

Block of responses to high-efficacy agonists by this sulfated steroid is greater than block of responses to partial agonists at saturating concentrations (Eisenman et al., 2003; Wang et al., 2002). This is called “activation dependant” or “state dependant inhibition”. Picrotoxinin, another GABA channel blocker superficially similar to pregnenolone sulfate in its activation dependence but the site of action of pregnenolone sulfate does not require a functional picrotoxin site for inhibition of GABA responses (Shen et al., 1999). GABA receptor antagonist Zn2+ also acts in activation dependant manner. However, Zn2+ binding site is located in the interface between α- and β-subunits and is activity is minimal in ternary αβγ receptor (Hosie et al., 2003).

6.2 3β-hydroxy steroids 6.2.1 Mode of action of 3β-hydroxy steroids Considerable effort has been given to characterize the effect of 3β- hydroxysteroids. For potentiating activity, OH-group at 3α-stereochemistry is essential regardless of the configuration at C5. An advantage of incorporating 3β- substituents is that potential hormonal activity resulting from oxidative

37 metabolism is abolished, and oral bioavailability is improved (Hogenkamp et al., 1997). Several earlier studies explored 3β-steroids as potential steroid antagonists. [3H]Flunitrazepam binding assays suggest that 3β-hydroxysteroids competitively antagonize the actions of the potentiating steroids (3α,5α)-3-hydroxypregnan-20- one (3α5αP) and (3α,5β)-3-hydroxypregnan-20-one (3α5βP), two standard representatives of the 5α-reduced and 5β-reduced classes of potentiating steroids (Prince and Simmonds, 1992, 1993). In electrophysiological studies, 3β5βP antagonized the 3α5βP-induced enhancement of GABA current (Maione et al., 1992; Maitra and Reynolds, 1998). However, only 3β5αP diminished the inhibitory effects of 3α5αP and 3α5βP on population spikes evoked in rat hippocampal CA1 stratum pyramidal (Wang et al., 2000). Studies on the chloride uptake into synaptosomes in rat cerebral cortex, in hippocampus and sIPSP in medial preoptic nucleus (MPN) showed that 3β-steroids reduced the 3α5αP-enhanced GABA response (Stromberg et al., 2006). In the presence of only GABA and absence of 3α5αP, some of the 3β-steroids potentiated GABA-evoked chloride ion uptake and prolonged the spices decay time, whereas the others had little or no effect on GABA stimulated current (Stromberg et al., 2006).

6.2.2 Modification of 3β-position Along with α in C5, 3β-ethyl, -propyl, -trifluoromethyl and -(benzyloxy)methyl, as well as substituents of the form 3β-XCH2, where X is Cl, Br, or I or contains unsaturation, show limited efficacy in inhibiting [35S]TBPS binding. An inhibition 35 of [ S]TBPS binding indirectly indicates an agonistic action on the GABAA receptor. With 5β, all of the 3β-substituted derivatives of pregnanolone inhibit TBPS via a single class of binding sites. In addition, all of the 3-substituted 5β- sterols tested are full inhibitors of [35S] TBPS binding (Hogenkamp et al., 1997). Electrophysiological measurements using α1β2γ2L receptors expressed in oocytes showed that 3β-methyl- and 3β-(azidomethyl)-3α-hydroxy-5α-pregnan-20-one are potent full efficacy modulators and that 3α-hydroxy-3β-(trifluoromethyl)-5α- pregnan-20-one is a low-efficacy modulator (Hogenkamp et al., 1997). These results indicate that protection of the OH-group at the 3α position by modification of the 3β-position in 3α5αP and 3α5βP maintains activity at the neuroactive steroid site on the GABAA receptor. In animal studies, (CCD 1042) is an orally active anticonvulsant steroid, while the naturally occurring progesterone metabolites 3α5αP and 3α5βP are inactive when administered orally, suggesting that 3β-substitution slows metabolism of the 3-hydroxyl, resulting in orally bioavailable steroid modulators of the GABAA receptor.

Moreover, another substitution at 3β position; 3β-(p-acetylphenylethynyl)-3α- hydroxy-5α-pregnan-20-one (Co 152791) was more potent than 3α5αP both in [35S]TBPS binding in human recombinant receptor and Xenopus oocyte expressing

38 GABAA receptor. Ganaxolone (CCD 1042), another 3β-methyl-substituted analog of the endogenous neuroactive steroid 3α5αP has a high-affinity, stereoselective, as positive allosteric modulator of the GABAA receptor complex and exhibits potent anticonvulsant activity across a range of animal procedures. The profile of anticonvulsant activity obtained for ganaxolone supports clinical evaluation of this drug as an antiepileptic therapy with potential utility in the treatment of generalized absence seizures as well as simple and complex partial seizures (Carter et al., 1997).

Screening for the selective antagonists of neuroactive steroid potentiation of GABA responses is going on in this lab and many other labs. Some success has been achieved in this regard. An steroid analogue 3α,5α-17-phenylandrost-16-en- 3-ol(17PA) blocked the 3α5αP-enhanced GABA response and direct GABA- mimetic effect of 3α5αP (Mennerick et al., 2004). However, 17PA had no or little effect on 3α5βP steroid. The main drawback of this substance is that the results cannot be reproduced because of difficulty to dissolve and there is no effect up to 5 µM concentration. Furthermore, this substance demands further evaluation in the animal behavior model.

6.2.3 Mechanism of action of 3β-hydroxy steroids How 3β-hydroxy steroids act on GABAA receptor is not fully understood. Although 3β-hydroxysteroids reduced the potentiation induced by 3α- hydroxysteroids, 3β-hydroxysteroids act noncompetitively with respect to potentiating steroids and inhibited the largest degrees of potentiation most effectively. The profile to block activation dependant inhibition was similar to that exhibited by sulfated steroids, known antagonist of GABAA receptors (Wang et al., 2002). These antagonist neuroactive steroids are also direct, noncompetitive GABAA receptor antagonists in recombinant receptors expressed in Xenopus oocytes but this was not seen when chloride uptake was studied using synaptosomes from rat cortical tissue (Lundgren et al 2003). 3β-hydroxysteroids also inhibit the potentiating effect of barbiturates and flunitrazepam in the oocyte model. Therefore, 3β-hydroxypregnane steroids are not direct antagonists of potentiating steroids but rather are noncompetitive, likely state-dependent, blockers of GABAA receptors (Wang et al., 2002).

3β-hydroxysteroids and pregnenolone sulfate-like GABAA receptor antagonists block GABAA receptors more effectively under conditions with open channels suggesting that the antagonism of GABA responses may represent ligand- dependent or state-dependent block. This direct non-competitive effect of 3β- hydroxysteroids and pregnenolone sulfate on the GABA-response was sufficient to account for the apparent antagonism of agonist steroids. However, if the effect

39 of 3β-hydroxysteroids and pregnenolone sulfate share the common binding property is still unknown. Because of the potential importance of antagonist steroids as experimental and clinical use, it is essential to characterize the functional and binding properties of 3β-hydroxysteroids and pregnenolone sulfate.

7. Recombinant GABA receptor: Advantages and disadvantages with studies on recombinant receptors expressed in oocyte over cell lines Although other expression systems like cell line have been developed, certain experiments are only feasible in oocytes. The expression of foreign proteins in Xenopus oocytes has many advantages for electrophysiological measurements. Numerous combinations of GABA as well as other receptor subunits can be expressed in the Xenopus oocyte. It is therefore possible to obtain results from several subunit combinations and including mutated receptor subunits. The size of ooctyes is large (~1 mm) which easily accommodates manipulations like mRNA injections and electrode penetration is certainly one aspect. However, the fact that it is possible to obtain oocyte attached patch clamp recordings (Methfessel et al., 1986) and, in particular, recordings from macropatches (Stuhmer et al., 1987) is among the main advantages. The low noise recordings from a large number of channels would be difficult to obtain from other systems, and measurements, like gating current fluctuations (Conti and Stuhmer, 1989) are presently limited to the oocyte expression system.

There are some drawbacks of the oocyte expression system. There is uncertainty in homogenous results in the use of different batches of oocytes and different preparations of mRNA. These factors contribute to the diversity of results, especially when one tries to quantify them. The oocytes from different donors may differ significantly in the levels of internal second messengers and in resting membrane properties; both factors may affect the properties of the exogenous channels (Dascal, 1987). This is an inevitable disadvantage of the method, and one has to aware of it when trying to draw any generalized conclusions. The only way to avoid mistake that may be introduced by such variability is to compare as many batches of oocytes and RNA as possible. Another disadvantage is that the size to the oocyte hinders rapid applications of ligands to the whole oocyte and makes it difficult to use for the kinetics of the receptor function.

The mammalian cell expression systems also offer many advantages over other expression systems such as Xenopus oocytes. For instance, a large number of the transfected mammalian cells can be generated and passaged by standard culture techniques. Stable transfection is feasible in mammalian cells as opposed to Xenopus oocytes that can only be transiently transfected. Stable transfection in cell

40 clones shows fewer variations between experiments and batches of cells. However, to study the recombinant combinations of different GABAA receptor in a shorter time frame as it was done in these current pharmacological studies, Xenopus oocytes is the better choice.

8. Uses of rodent models for the testing of drugs for preclinical evaluation To date, the behavioral and neurological studies on effects of active GABAA receptor modulators, for example neuroactive steroids, are performed in rat models for human disorders (Johansson et al., 2002; Lofgren et al., 2006; Miczek et al., 2003; Smith et al., 1998; Turkmen et al., 2004). It is a general concept that the response of neuroactive steroids in human would be similar as in rat. There are also arguments against this concept. As the GABAA receptor modulation is dependant on the subunit combination, the pharmacology and steroid modulation of the mammalian GABAA receptor might vary between two species. It is therefore essential to investigate differences in receptor pharmacology and steroid modulation between the human and rat GABAA receptor. There are few evidences showing the difference of receptor pharmacology between the human and rat species. In a previous study by Nguyen and co-worker 1995 showed that modulation of GABAA receptor binding site in human brain by neuroactive steroids and the effects compared to those in rat brain. Unlike rat, human brain binding of [3H]flunitrazepam to the benzodiazepine site was not enhanced by alphaxalone (at any concentration), but was unaffected in many regions and inhibited in others. Binding of [3H]muscimol to high and low affinity GABA sites were enhanced by both steroids in all tested regions of rat brain (Nguyen et al., 1995). Another study showed that the tonic conductance in rat cerebellar granule cells (CGS), in contrast to that in mouse CGSs, is relatively insensitive to 3α5αTHDOC (Hamann et al., 2002). Species differences are consistent with the existence of GABAA receptor subtypes that differ in pharmacology. On the other hand, the differences in GABA receptor properties between human and rat brain, not due to the major differences in the gene sequence (Tyndale et al., 1994), but apparently in the ratios of the subunit expressed in a given brain area as indicated by binding heterogeneity (Nguyen et al., 1995). This piece of evidence suggests caution in use of rodent models for the testing of drugs such as antiepileptics or anxiolytics for use in human. Careful consideration of the similarities and differences between species may allow use of laboratory animals for certain types of preclinical research questions.

41 Aims: The overall goal of the present work was to study the effects of neuroactive metabolites of sex and stress hormones in recombinant GABAA receptor.

The specific aims of different papers were:

1. To investigate the pharmacological properties of α1-, α4- and α5-subunit containing GABAA receptors.

2. To study if the binding property of pregnenolone sulfate (PS) is different from that of Zn2+.

3. To study if the pharmacological property of 3β-hydroxysteroids is different from that of pregnenolone sulfate.

4. To study if the neuroactive steroids activity is different between the recombinant human and rat α1β2γ2L GABAA receptors.

5. To investigate the activity of neuroactive steroids between 2L (long form) and 2S (short from) variant of γ-subunit of human GABAA receptor.

42 Method Chemicals The composition of the standard ND96 solution was (in mM): 96 NaCl, 1 KCl, 1 MgCl2, 2 CaCl2 and 5 HEPES at pH 7.4. Pentobarbital was dissolved in 0.1 M NaOH as a stock solution of 25 mM, and subsequently diluted with ND96. Zn2+ was dissolved in water. The 3α-hydroxypregnane steroids used in this project were 3α-5α- androstane carbonitrile (3α5αACN); 3α-OH-5α-pregnan-20-one (allopregnanolone, 3α5αP); 3α-OH-5β-pregnan-20-one (pregnanolone, 3α5βP), 5α-pregnan-3α-ol-11,20-dione (Alphaxalone); 4-pregnan-3α-ol-20-one, 5α- androstane-3α,17β-diol (3α5α-ADL); 5α-pregnan-3α,21-diol-20-one (3α5αTHDOC); 5β-pregnan-3α,21-diol-20-one (3α5βTHDOC); 5α-pregnane- 3α,20β-diol and 5β-pregnane-3α,20α-diol. The 3β-hydroxypregnane steroids were 3β-OH-5α-pregnan-20-one (UC1010), 5α- pregnan-3β, 20β-diol (UC1011), 5β-pregnan-3β, 20α-diol (UC1013), 3β-OH-5β- pregnan-20-one (UC1014), 5β-pregnan-3β, 21-diol-20-one (UC1015), 5α-pregnan- 3β, 20α-diol (UC1019) and 5β-pregnan-3β, 20β-diol (UC1020). Steroids were dissolved in dimethyl sulfoxide (DMSO) except pregnenolone sulfate which was dissolved in the distilled water. All chemicals were then diluted in external solution for experiments. We tested the effect of DMSO in GABAA receptor expressed in Xenopus oocytes and did not find any significant response of DMSO up to 0.2% in the α1β2γ2L receptor. However, the concentration of DMSO in experimental solutions was ≤ 0.1%. cDNA cloning The cDNAs encoding the rat GABAA-receptor subunits α1, β2 and γ2 were provided by A. Tobin, University of California, Los Angels, CA, USA (α1), P. Malherbe, Hoffman-La Roche, Switzerland (β2), C. Fraser, National Institute on Alcohol Abuse and Alcoholism, Bethesda, MD, USA (γ2L). The cDNAs encoding for rat α4, α5 and of all human subunits were prepared in this lab (for details see individual articles). These plasmids have been used successfully for in vitro transcription where mRNAs were injected into Xenopus oocytes, and produced receptors were examined with electrophysiology.

In vitro transcription and expression in the Xenopus oocyte Sexually mature female Xenopus laevis were kept in optimal condition by authorized animal keepers in a local animal house and fed with standard frog food. Experimental protocols were approved by the Animal Experimentation Ethical Committee in Umeå. Stage V-VI oocytes were harvested under 0.1% tricaine (3- aminobenzoic acid ethyl ester) anesthesia. mRNA (20-40 ng total RNA) was injected 24 hours following defolliculation. Initial studies on oocytes injected with

43

Figure 10. Female Xenopus laevis frog and its oocyte

α5, β2 and γ2L mRNA molar ratio of 1:1:1 revealed a variable level of GABA EC50s and Imaxs after repeated tests. Comparatively, stable and consistent EC50s and Imaxs were observed in 2:1:1 and 3:1:1 ratios in three repeated (5~7 cells in each) experiments. On the contrary, α1, β2, and γ2L subunit mRNAs injected in equimolar ratio (1:1:1) showed a stable GABA EC50s and average Imaxs after repeated tests. Therefore, the experiments in these studies were performed at molar ratio of 2:1:1 in the α5β2γ2L and molar ratio of 1:1:1 in other rat wild-types and mutant receptors as well as human receptor.

A Nanoject II auto nanoliter injector (Drummond Sci. co., PA, USA) was used for mRNA microinjection. Oocytes were incubated up to 5 days at 18 °C in the standard ND96 at pH 7.4, supplemented with pyruvate (5 mM), (100 U/ml), streptomycin (100 µg/ml) and gentamycin (50 µg/ml).

Oocyte electrophysiology

Figure 11. Oocyte electrophysiology with voltage clamp technique

Two-electrode voltage-clamp whole-cell recordings were performed with a Warner OC725 amplifier (Warner instrument corporation, Hamden, CT, USA) 2-3 days following RNA injection. The extracellular recording solution was ND96 medium with no supplements. Intracellular recording pipettes were filled with 3 M KCl and

44 had open tip resistances of ~ 1 MΩ. Drugs were bath applied from a common tip via a gravity-driven multibarrel drug-delivery system-ValveLink 16 pinch valve perfusion system which was controlled by a Valvelink 16 controller. Cells were clamped at -70 mV for all experiments, and the steady-state current at the end of 20s drug applications was measured for quantification of current amplitudes. At the steady-state, the open and close state of Cl- channels reached equilibrium. We measure the response at the end part of drug application because of relatively slow rate of drug application. In this stage the response of lower concentrations gives a relatively horizontal and linear response and the peak and steady response are almost identical in height. However, peak response is relatively higher than steady- state response at higher concentration. The interval between the drug applications was 90 s.

Data analysis Data acquisition and analysis were performed with pClamp software (Axon Instruments, San Francisco, CA, USA). Data plotting and curve fitting were done with Sigma Plot software (SPSS, Chicago, IL, USA). Modulation of GABA- activated current was calculated as I = (IM - IN)/IN, where IN (Normalizing current) and IM (measured current) are the amplitudes of the chloride current in the absence and presence of the test drugs, respectively. Fitting of the dose-response relationships was performed using the Hill equation.

Statistical analysis Unless otherwise specified, statistical differences were determined using a two- tailed Student's t test. Analysis of variance (ANOVA) was used to analyze the effect of 3α and 3β-steroids between the human and rat α1β2γ2 receptors. With significant results (p< 0.05), Least Significant Difference (LSD) post-hoc test was used. Repeated measures analysis of variance (ANOVA) was also used to compare the efficacies of neuroactive steroids within each receptor (p< 0.05). Dose- response curve were fitted to average curve values using the Hill equation as n n n follows: I = Imax × {C /(EC50 + C )}, where C is the concentration of drug, Imax is the maximum current amplitude, EC50 is the concentration of drug that produces

50% of Imax, and n is the Hill coefficient.

All the values in the text and figures as mean ± S.E.M. Significance level used was p<0.05. The SPSS statistical package (SPSS, Chicago, IL, USA) version 12 was used for data analysis. For plotting of curves SigmaPlot 8.0 was used.

45 Results

GABA site antagonism and pentobarbital inhibition by bicuculline is not dependant on α-subunit (Paper I) The effect of bicuculline methiodide (BMI), an antagonist in the GABAA receptor binding site was compared in recombinant GABAA receptors containing α1-, α4- and α5-subunits. We found that the dose-inhibition curve of bicuculline to inhibit the sub-maximal (30 µM) GABA response was identical with the α1β2γ2L, α4β2γ2L and α5β2γ2L receptors (Paper I, figure 2a). Inhibition of the pentobarbital-activated direct currents by bicuculline was also observed in above mentioned receptor combinations. Likewise, the potencies of bicuculline to inhibit sub-maximal (1 mM) pentobarbital response were equivalent with the α1β2γ2L, α4β2γ2L and α5β2γ2L receptors (Paper I, figure 2b). The efficacies of bicuculline in terms of maximal inhibition of the 1 mM pentobarbital response were also identical (about 65% from control) in these receptors. Moreover, the presence of a fixed concentration (10, 30 or 100 µM) of bicuculline shifted the GABA-response curve to the right without decreasing the maximum response at the α1-, α4- and α5-subunit containing receptors. This type of inhibition is called competitive antagonism. On the other hand, the bicuculline inhibited pentobarbital activated current in a non-competitive manner; depression of maximum response with no significant change in EC50 (Paper I, figure 3).

Pregnenolone sulfate response in not dependant on γ-subunit (Paper II) Pregnenolone sulfate, like 3β-OH steroids, is a potent inhibitor of GABA- mediated and 3α5αTHDOC-mediated response. The dose-inhibition curves of PS to inhibit GABA and to inhibit pentobarbital response at the α1β2γ2L and α1β2 receptor combination were observed. The efficacy and potency of pregnenolone sulfate to block equivalent concentration of GABA or pentobarbital was identical at the ternary α1β2γ2L and binary α1β2 receptor (Paper II, Figure 3a and 3b). This result indicated that the γ2-subunit did not affect the potency of PS to inhibit GABAA receptor-mediated currents.

Pregnenolone sulfate and Zn2+ do not share the same binding site (Paper II) The inhibitory effect of pregnenolone sulfate was characterized in two mutant receptors- α1V256Sβ2γ2L and α1β2A252Sγ2L. In these two mutants, valine and residues at 2’position closest to the cytoplasmic end of the M2 helix were replaced by serine at α1 and β2 subunit, respectively. In both of these mutant receptors the inhibitory effect of pregnenolone sulfate to inhibit GABA was

46 significantly reduced (Paper II, figure 4e and 4f). Unlike PS, Zn2+ inhibition was not reduced at the mutant α1V256Sβ2γ2L and α1β2A252Sγ2L receptors and their degree of inhibition was similar to that with the wild-type α1β2γ2L receptor (Paper II, figure 4d). This indicates that pregnenolone sulfate does not share the same binding site of Zn2+ and that PS effects is influenced by the amino acid changes at the mutation sites.

Functional differences between 3β-hydroxysteroids and pregnenolone sulphate (Paper III) Homologous mutation of the residue at 2’position closest to the cytoplasmic end of the M2 helix to serine on both α1 and β2 subunit, α1V256S and β2A252S, reduced the slow desensitization components of GABA-activated currents at saturating doses. Compared to the wild type receptor, the potency of GABA increased significantly at the α1V256Sβ2γ2L receptor (P < 0.05), whereas it decreased moderately at the α1β2A252Sγ2L receptor (paper III, Fig. 2). We found that PS, 5α-pregnan-3β, 20(S)-diol (UC1019) and β-pregnan-3β, 20(R)-diol (UC1020) were potent antagonists at the wild type receptor. The inhibitory effect of PS was significantly reduced at both mutant α1V256Sβ2γ2L and α1β2A252Sγ2L receptors (P < 0.05), whereas the inhibitory effects of UC1019 and UC1020 were reduced only at the mutant α1V256Sβ2γ2L receptor (paper III, Fig. 3) but not at the α1β2A252Sγ2L receptors. PS promoted slow desensitization with prolonged GABA application in a dose-dependent manner at the wild type receptor, but not at the mutant receptors. On the contrary, UC1019 and UC1020 (≤ 20 µM) did not promote desensitization with both wild type and mutants receptors.

The neuroactive steroids activity is different between the recombinant human and rat α1β2γ2L GABAA receptors (Paper IV) The effect of agonist (3α-OH) or antagonist (3β-OH) neuroactive steroid series was also investigated at human and rat α1β2γ2L to find if they show any difference between the species. Both the potencies and efficacies of GABA were significantly higher at the human α1β2γ2L receptor. Enhancement of GABA EC15 response by 3α-OH-5β-pregnan-20-one (3α5βP), 5α-androstane-3α,17β-diol (3α5αADL) and 5α-pregnane-3α,20β-diol (3α5α-diol) was significantly higher at the rat receptor compare to the human receptor (paper IV, figure 3, Table 2). 3α5αP at the human and ∆4-pregnen-3α-ol-20-one at the rat receptor are more potent among agonist steroids. Additionally, diol-containing steroids showed lower degree of potentiation in both receptors. The antagonistic effects of 5α- pregnan-3β,20β-diol (UC1011) and 5α-pregnan-3β, 20(S)-diol (UC1019) at the

47 human receptor were significantly higher to inhibit the 3α5αTHDOC + GABA response compared to the rat receptor and pregnenolone sulfate in the rat receptor compared to the human receptor (paper IV, Fig. 4). Inhibition of 30 µM GABA response by 3β-OH-5α-pregnan-20-one (UC1010) and 5β-pregnan-3β, 21-diol-20- one (UC1015) was significantly higher at the human receptor compared to that in the rat and 5β-pregnan-3β, 20(R)-diol (UC1020) at the rat receptor compared to the human receptor (P <0.05).

Neuroactive steroids activity differs between 2L (long form) and 2S (short from) variant of γ-subunit of human GABAA receptor (Paper IV) Potentiation by of 3α5αTHDOC (0.1~3 µM) and 3α5βP (0.1~1 µM) was significantly higher with the γ2S isoform compared to the γ2L isoform (Paper IV, Figure 5). Moreover, inhibition of 30 µM GABA by UC1010 was significantly higher with the receptor containing γ2L and UC1020 in the receptor containing γ2S isoform, respectively. To inhibit 3α5αTHDOC + GABA response, effects of UC1010 and UC1019 were significantly higher with the γ2L isoform (P <0.05 Paper IV, Figure 5). However, GABA dose-response curves were identical for human GABAA receptors with γ2S and γ2L subunits except for an increased maximum GABA-evoked current at the γ2L subunits. Conversely, the direct effect of pentobarbital was significantly higher with the γ2S isoform.

Discussion

GABA and pentobarbital inhibition by bicuculline at α1-, α4-, α5- subunit containing GABAA receptor Our results showed that neither GABA site antagonism nor the pentobarbital inhibition by bicuculline is dependant of α-subunit. GABA-site inhibition by bicuculline at the α1-containg receptors is called phasic inhibition and inhibition at the α4- and α5-contaning extrasynaptic receptors is called tonic inhibition. It was reported earlier that the tonic inhibition observed at the α4- and α5-containing GABAA receptors was known to be more resistant to the GABAA receptor antagonists compared to the phasic inhibition in the α1-containing receptors (Farrant and Nusser, 2005; Stell and Mody, 2002). We did not find any functional difference of bicuculline between recombinant rat GABAA receptors mediating phasic and tonic inhibition since high concentrations of bicuculline in the micromolar range were tested in the present study. At submicromolar concentrations, several competitive and non-competitive GABAA antagonists (gabazine, picrotoxin and bicuculline) reduced phasic currents, but had no effect

48 on tonic currents. Moreover, the antagonists blocked both phasic and tonic currents at high concentrations (Bai et al., 2001; Semyanov et al., 2003; Stell and Mody, 2002; Yeung et al., 2003). However, low submicromolar concentrations were not tested in the present study. In our study, pentobarbital was a direct activator of α1-, α4- and α5-containing receptor and bicuculline inhibited the pentobarbital-activated currents in a non- competitive manner. However, another report demonstrated a lack of direct activation of the recombinant human α4β1γ2 receptor by pentobarbital in the oocyte expressing system (Wafford et al., 1996). On the other hand, in agreement with our findings, direct activation of α4-containing receptors by pentobarbital in the present study is with a previous study with direct activation of human α4β2γ2L receptors expressed in Xenopus oocyte by pentobarbital (Whittemore et al., 1996). Our studies also showed that the effect of GABA is dependent on γ-subunit. The potencies of GABA at the α1β2, α4β2 and α5β2 binary receptors were significantly higher compared to the ternary α1β2γ2L, α4β2γ2L and α5β2γ2L receptors. The efficacy of GABA currents evoked at the binary α1β2 and α4β2 receptors were equivalent to those evoked at the ternary α1β2γ2L and α4β2γ2L receptors. However, the α5β2 receptor possessed significantly lower GABA efficacy compared to the α5β2γ2L receptor. This indicates that for proper sorting, assembly and expression three subunits are essential for α5-containing receptor (Burgard et al., 1996).

Pregnenolone sulfate response is not dependant on γ-subunit The effect of pregnenolone sulfate at the binary α1β2 in this study is not significantly different from ternary α1β2γ2L receptor. Therefore, the inhibition by pregnenolone sulfate was not primarily dependent on the γ2-subunit. The presence of serine at the 2’ residue N-terminal to the beginning of M2 helix at both mutant α1 and β2-subunit significantly reduced PS inhibition on the GABAA receptor-mediated current. Amino acid residues underlying the inhibition of Cl- channel by PS may locate at channel lining segments linked to V256 at the α1 and A252 at the β2 subunits. Zn2+ is also a potent inhibitor of GABA binding site but its efficacy dependant on the absence of γ subunit. Mutations of amino acid residues in the above positions at the α1 and β2 subunits did not abolish the inhibitory effect of Zn2+ indicating that Zn2+ dose not share the same binding with pregnenolone sulfate. Moreover, an earlier study showed that Zn2+ has four binding site; one is located within the ion channel, other two are on the external amino (N)-terminal face of the receptor at the interfaces between alpha (E137 & H141) and beta (E182) subunits (Hosie et al., 2003). A recent study showed that four residues in the first transmembrane domain are required for the majority of the sensitivity to PS, but a charged extracellular residue at the end of the M2 helix

49 also plays a role (Wardell et al., 2006). These residues are also different from the residues responsible for Zn2+ binding.

3β-hydroxysteroids and pregnenolone sulfate inhibit recombinant rat GABAA receptor through different mechanisms Pharmacological analysis of antagonist steroids in the recombinant GABAA receptors has provided insight about functional differences between 3β- hydroxysteroids and pregnenolone sulfate. Our results suggest that the inhibition by PS, but not in the same degree by 3β-hydroxysteroids, on the GABAA receptor was dependent on the receptor desensitization. We further showed that a point mutation at M2 helix of the β2-subunit (β2A252S) can reduce the inhibitory effect of PS on the GABAA receptor without affecting the inhibitory properties of 3β- hydroxysteroids. Given that both PS and 3β-hydroxysteroids inhibit GABAA receptor-mediated currents in an activation-dependent manner, it is possible that PS inhibits GABAA receptor function by promoting receptor desensitization, whereas the inhibitory properties of 3β-hydroxysteroids are more based on other channel properties of the GABAA receptor.

The neuroactive steroids activity between the human and rat α1β2γ2L recombinant GABAA receptors Some differential response agonist and antagonist neuroactive steroids have been observed between the human and rat α1β2γ2L receptor. Some steroids showed more effect in the human receptor than rat receptor or vice -versa. Although very few reports are available regarding the comparative study between the species, one report corresponds to our finding that the neuroactive steroids effects vary between human and rat brain (Nguyen et al., 1995). Modulation GABAA receptor binding site by 3α5αTHDOC and alphaxalone varied between the human and rat brain in that study. The effects of neuroactive steroids in human disorder have been evaluated in the animal behavioral model (Johansson et al., 2002; Lofgren et al., 2006; Turkmen et al., 2004) and it is generally believed that the effects of neuroactive steroids on the GABAA receptor are also similar in human. Our results along with the above report suggest that the effect of neuroactive steroids in the rat model might be different from that in human. This heterogeneity between species provides the information that testing of potential should be made in human receptors rather than rat receptors. That may give insights to how to interpret the results from animal models and shows that result in animal studies should take into account differences between species. Our findings also suggest that further studies are clearly needed to find out any similarity and difference of other pharmacological agents between human and rat GABAA receptor.

50 Differential response of neuroactive steroids activity between 2L (long form) and 2S (short from) variant of γ-subunit of human GABAA receptor Our results show some difference of neuroactive steroid modulation between 2L (long form) and 2S (short from) of γ-subunit. This may be due to the part missing in the γ2S form of γ-subunit. γ2L differ from γ2S by having an additional 24 nucleotide exons which encodes an octapeptide (Whiting et al., 1990). The octapeptide includes a consensus phosphorylation site for PKC, with a serine in position 343. Inclusion or exclusion of this octapeptide and its phosphorylation site may alter the phosphorylation of resulting receptor and lead to altered receptor function (Kellenberger et al., 1992; Lin et al., 1996; Moss et al., 1992). Therefore, it was predicted that there would be difference in the neuroactive steroids effect between these two variant of γ-subunit. It was also reported that prolong application of some therapeutic agents can change the γ2L/S mRNA ratios. Prolong application of pentobarbital reduces the ratio. However, it was evident that 3α5αP has a little effect on the γ2L/S ratio (Tyndale et al., 1997).

51 Conclusions

GABA site antagonism and pentobarbital inhibition by bicuculline is not dependant on α-subunit. GABA potency (not efficacy) is dependant on γ-subunit. Pregnenolone sulfate (PS) response is not dependant on γ subunit and PS and 2+ Zn acts on GABAA receptor through two separate mechanisms.

3β-hydroxysteroids and pregnenolone sulfate inhibit recombinant rat GABAA receptor through different mechanisms. The neuroactive steroids activity is different between the recombinant human and rat α1β2γ2L GABAA receptors. Neuroactive steroids activity differs between 2L (long form) and 2S (short from) variant of γ-subunit of human GABAA receptor.

52 Acknowledgement

I would like to express my sincere gratitude to all the people who have supported me throughout this work. In particular, I would like to thank:

Associate professor Ming-De Wang, my supervisor, who with patience, scientific skill and generosity guided me into the world of neuroscience. Thank you very much for giving me time in spite of your intense clinical works. Without your help I could not come to this point.

Professor Torbjörn Bäckström, my co-supervisor for accepting me as a PhD student in his group. With a never-ending enthusiasm, encouragement and pedagogical skill you have helped me a lot to acquire scientific works. Thank you very much for your kindness and warm heart.

Associate professor Inga-Maj Johansson, thank you for your generous ideas, valuable comments and discussion and proof-reading of my manuscripts.

Monica Isaksson, Every time I needed the mRNA, you made that for me. It was so kind of you to show me all the technical steps of cloning and producing mRNAs. Professor Ann Lalos, head of the Department of Obstetrics and Gynecology for her friendly support. Dr. Gianna Ragagnin, thank you for the fruitful discussions about the chemical structures of steroids and friendship and of course for ‘tiramisu’. Associate professor Per Lundgren, thank you for your cooperation in the lab and help me to find references. Dr. David Haage, thank your for your fruitful discussion about ion channels. Elizabeth Zingmark and Agneta Andersson, for all the help in the laboratory assistance. Dr. Magdalena Taube, thank you very much for the friendship and keeping our lab environment so live. Jessica Stromberg and Charlotte Lindblad, for friendship and assistance in laboratory and personal issues from the beginning of my study.

Magnus Löfgren, we always shared our feeling each other and had common views in many worldly matters. Tobias Näslund, for friendship and assistance during your stay at the UNC. Birgit Ericson and Carina Jonsson, for your kind administrative support.

53 My friends and present and former co-workers at UNC, Di Zhu, Sahruh Turkmen, Vita Birzniece, Sigrid Nyberg, Lotta Andreen, Erika Timby, Anna-Carin Wihlbäck and all the fantastic people at the department of obstetrics and gynecology. Marianne Borgström and Lena Gustavsson, for taking good care of the frogs and for assistance. Finally to my wife Tanzina Chowdhury and kid Yazdan Rahman (Evan), for your endless support throughout my whole life. This work would have not been finished without your love and understanding.

This study was done at the Umeå Research Centre (UNC) under the Department of Clinical Science, Obstetrics and Gynecology, Umeå University and supported by the Swedish Research Council-medicine (project No. 4X-11198; 73p-15450), Insamlingsstiftelsen för medicinsk forskning vid Umeå Universitet, Svenska läkaresällskapet, Socialstyrelsens forskningsfond, Umeå Kvinnoklinikens forskningsfond and the EU regional fund’s objective 1 program.

54 References

Adkins, C.E., Pillai, G.V., Kerby, J., Bonnert, T.P., Haldon, C., McKernan, R.M., Gonzalez, J.E., Oades, K., Whiting, P.J., Simpson, P.B., 2001. alpha4beta3delta GABA(A) receptors characterized by fluorescence resonance energy transfer-derived measurements of membrane potential. J Biol Chem 276, 38934-38939. Akk, G., Bracamontes, J., Steinbach, J.H., 2001. Pregnenolone sulfate block of GABA(A) receptors: mechanism and involvement of a residue in the M2 region of the alpha subunit. J Physiol 532, 673-684. Akk, G., Bracamontes, J., Steinbach, J.H., 2004. Activation of GABA(A) receptors containing the alpha4 subunit by GABA and pentobarbital. J Physiol 556, 387-399. Akk, G., Shu, H.J., Wang, C., Steinbach, J.H., Zorumski, C.F., Covey, D.F., Mennerick, S., 2005. Neurosteroid access to the GABAA receptor. J Neurosci 25, 11605-11613. Alves, S.E., Weiland, N.G., Hayashi, S., McEwen, B.S., 1998. Immunocytochemical localization of nuclear estrogen receptors and progestin receptors within the rat dorsal raphe nucleus. J Comp Neurol 391, 322-334. Backstrom, T., 1976. Epilepsy in women. Oestrogen and progesterone plasma levels. Experientia 32, 248-249. Backstrom, T., Andersson, A., Andree, L., Birzniece, V., Bixo, M., Bjorn, I., Haage, D., Isaksson, M., Johansson, I.M., Lindblad, C., Lundgren, P., Nyberg, S., Odmark, I.S., Stromberg, J., Sundstrom- Poromaa, I., Turkmen, S., Wahlstrom, G., Wang, M., Wihlback, A.C., Zhu, D., Zingmark, E., 2003. Pathogenesis in menstrual cycle-linked CNS disorders. Ann N Y Acad Sci 1007, 42-53. Backstrom, T., Andreen, L., Birzniece, V., Bjorn, I., Johansson, I.M., Nordenstam-Haghjo, M., Nyberg, S., Sundstrom-Poromaa, I., Wahlstrom, G., Wang, M., Zhu, D., 2003. The role of hormones and hormonal treatments in premenstrual syndrome. CNS Drugs 17, 325-342. Backstrom, T., Gee, K.W., Lan, N., Sorensen, M., Wahlstrom, G., 1990. Steroids in relation to epilepsy and anaesthesia. Ciba Found Symp 153, 225-230; discussion 230-229. Baer, K., Essrich, C., Benson, J.A., Benke, D., Bluethmann, H., Fritschy, J.M., Luscher, B., 1999. Postsynaptic clustering of gamma-aminobutyric acid type A receptors by the gamma3 subunit in vivo. Proc Natl Acad Sci U S A 96, 12860-12865. Bai, D., Zhu, G., Pennefather, P., Jackson, M.F., MacDonald, J.F., Orser, B.A., 2001. Distinct functional and pharmacological properties of tonic and quantal inhibitory postsynaptic currents mediated by gamma-aminobutyric acid(A) receptors in hippocampal neurons. Mol Pharmacol 59, 814- 824. Banerjee, P.K., Snead, O.C., 3rd, 1998. Neuroactive steroids exacerbate gamma-hydroxybutyric acid- induced absence seizures in rats. Eur J Pharmacol 359, 41-48. Barbaccia, M.L., Concas, A., Serra, M., Biggio, G., 1998. Stress and neurosteroids in adult and aged rats. Exp Gerontol 33, 697-712. Baulieu, E.E., 1998. Neurosteroids: a novel function of the brain. Psychoneuroendocrinology 23, 963- 987. Baulieu, E.E., Robel, P., 1995. Non-genomic mechanisms of action of steroid hormones. Ciba Found Symp 191, 24-37; discussion 37-42. Beauchamp, M.H., Ormerod, B.K., Jhamandas, K., Boegman, R.J., Beninger, R.J., 2000. Neurosteroids and reward: allopregnanolone produces a conditioned place aversion in rats. Pharmacol Biochem Behav 67, 29-35. Belelli, D., Casula, A., Ling, A., Lambert, J.J., 2002. The influence of subunit composition on the interaction of neurosteroids with GABA(A) receptors. Neuropharmacology 43, 651-661. Belelli, D., Herd, M.B., 2003. The contraceptive agent Provera enhances GABA(A) receptor-mediated inhibitory neurotransmission in the rat hippocampus: evidence for endogenous neurosteroids? J Neurosci 23, 10013-10020.

55 Belelli, D., Herd, M.B., Mitchell, E.A., Peden, D.R., Vardy, A.W., Gentet, L., Lambert, J.J., 2006. Neuroactive steroids and inhibitory neurotransmission: mechanisms of action and physiological relevance. Neuroscience 138, 821-829. Belelli, D., Lambert, J.J., 2005. Neurosteroids: endogenous regulators of the GABA(A) receptor. Nat Rev Neurosci 6, 565-575. Belelli, D., Lambert, J.J., Peters, J.A., Gee, K.W., Lan, N.C., 1996. Modulation of human recombinant GABAA receptors by . Neuropharmacology 35, 1223-1231. Belelli, D., Pau, D., Cabras, G., Peters, J.A., Lambert, J.J., 1999. A single amino acid confers barbiturate sensitivity upon the GABA rho 1 receptor. Br J Pharmacol 127, 601-604. Benson, J.A., Low, K., Keist, R., Mohler, H., Rudolph, U., 1998. Pharmacology of recombinant gamma-aminobutyric acidA receptors rendered diazepam-insensitive by point-mutated alpha-subunits. FEBS Lett 431, 400-404. Bethea, C.L., 1993. Colocalization of progestin receptors with in raphe neurons of macaque. Neuroendocrinology 57, 1-6. Beyenburg, S., Stoffel-Wagner, B., Bauer, J., Watzka, M., Blumcke, I., Bidlingmaier, F., Elger, C.E., 2001. Neuroactive steroids and seizure susceptibility. Epilepsy Res 44, 141-153. Birnir, B., Tierney, M.L., Dalziel, J.E., Cox, G.B., Gage, P.W., 1997. A structural determinant of desensitization and by pentobarbitone of the GABAA receptor. J Membr Biol 155, 157-166. Biswas, M.G., Russell, D.W., 1997. Expression cloning and characterization of oxidative 17beta- and 3alpha-hydroxysteroid dehydrogenases from rat and human prostate. J Biol Chem 272, 15959-15966. Bitran, D., Shiekh, M., McLeod, M., 1995. Anxiolytic effect of progesterone is mediated by the neurosteroid allopregnanolone at brain GABAA receptors. J Neuroendocrinol 7, 171-177. Bixo, M., Andersson, A., Winblad, B., Purdy, R.H., Backstrom, T., 1997. Progesterone, 5alpha- pregnane-3,20-dione and 3alpha-hydroxy-5alpha-pregnane-20-one in specific regions of the human female brain in different endocrine states. Brain Res 764, 173-178. Bixo, M., Backstrom, T., Winblad, B., Andersson, A., 1995. Estradiol and testosterone in specific regions of the human female brain in different endocrine states. J Steroid Biochem Mol Biol 55, 297- 303. Bjorn, I., Bixo, M., Nojd, K.S., Collberg, P., Nyberg, S., Sundstrom-Poromaa, I., Backstrom, T., 2002. The impact of different doses of acetate on mood symptoms in sequential hormonal therapy. Gynecol Endocrinol 16, 1-8. Blum, B.P., Mann, J.J., 2002. The GABAergic system in schizophrenia. Int J Neuropsychopharmacol 5, 159-179. Bogdanov, Y., Michels, G., Armstrong-Gold, C., Haydon, P.G., Lindstrom, J., Pangalos, M., Moss, S.J., 2006. Synaptic GABAA receptors are directly recruited from their extrasynaptic counterparts. Embo J 25, 4381-4389. Borghese, C.M., Storustovu, S., Ebert, B., Herd, M.B., Belelli, D., Lambert, J.J., Marshall, G., Wafford, K.A., Harris, R.A., 2006. The delta subunit of gamma-aminobutyric acid type A receptors does not confer sensitivity to low concentrations of ethanol. J Pharmacol Exp Ther 316, 1360-1368. Bormann, J., 2000. The 'ABC' of GABA receptors. Trends Pharmacol Sci 21, 16-19. Bormann, J., Feigenspan, A., 1995. GABAC receptors. Trends Neurosci 18, 515-519. Bowers, B.J., Wehner, J.M., 1992. Biochemical and behavioral effects of steroids on GABAA receptor function in long- and short-sleep mice. Brain Res Bull 29, 57-68. Brickley, S.G., Cull-Candy, S.G., Farrant, M., 1999. Single-channel properties of synaptic and extrasynaptic GABAA receptors suggest differential targeting of receptor subtypes. J Neurosci 19, 2960-2973. Brickley, S.G., Farrant, M., Swanson, G.T., Cull-Candy, S.G., 2001. CNQX increases GABA-mediated synaptic transmission in the cerebellum by an AMPA/kainate receptor-independent mechanism. Neuropharmacology 41, 730-736.

56 Brooks-Kayal, A.R., Shumate, M.D., Jin, H., Rikhter, T.Y., Kelly, M.E., Coulter, D.A., 2001. gamma- Aminobutyric acid(A) receptor subunit expression predicts functional changes in hippocampal dentate granule cells during postnatal development. J Neurochem 77, 1266-1278. Brussaard, A.B., Kits, K.S., Baker, R.E., Willems, W.P., Leyting-Vermeulen, J.W., Voorn, P., Smit, A.B., Bicknell, R.J., Herbison, A.E., 1997. Plasticity in fast synaptic inhibition of adult oxytocin neurons caused by switch in GABA(A) receptor subunit expression. Neuron 19, 1103-1114. Brussaard, A.B., Wossink, J., Lodder, J.C., Kits, K.S., 2000. Progesterone-metabolite prevents protein kinase C-dependent modulation of gamma-aminobutyric acid type A receptors in oxytocin neurons. Proc Natl Acad Sci U S A 97, 3625-3630. Brzozowski, A.M., Pike, A.C., Dauter, Z., Hubbard, R.E., Bonn, T., Engstrom, O., Ohman, L., Greene, G.L., Gustafsson, J.A., Carlquist, M., 1997. Molecular basis of agonism and antagonism in the oestrogen receptor. Nature 389, 753-758. Buhl, E.H., Otis, T.S., Mody, I., 1996. Zinc-induced collapse of augmented inhibition by GABA in a temporal lobe epilepsy model. Science 271, 369-373. Buhr, A., Baur, R., Malherbe, P., Sigel, E., 1996. Point mutations of the alpha 1 beta 2 gamma 2 gamma-aminobutyric acid(A) receptor affecting modulation of the channel by ligands of the benzodiazepine binding site. Mol Pharmacol 49, 1080-1084. Burden, P.M., Allan, R.D., Hambley, T., HJohnston, G.A., 1998. The synthesis of 2- cyclohexylideneperhydro-4,7-methanoindenes. Non-steroidal analogues of steroidal GABAA recepotr modulators. J. Chem. Soc. Perkin Trans. I, 3163-3167. Burgard, E.C., Tietz, E.I., Neelands, T.R., Macdonald, R.L., 1996. Properties of recombinant gamma- aminobutyric acid A receptor isoforms containing the alpha 5 subunit subtype. Mol Pharmacol 50, 119- 127. Callachan, H., Cottrell, G.A., Hather, N.Y., Lambert, J.J., Nooney, J.M., Peters, J.A., 1987. Modulation of the GABAA receptor by progesterone metabolites. Proc R Soc Lond B Biol Sci 231, 359-369. Calogero, A.E., Palumbo, M.A., Bosboom, A.M., Burrello, N., Ferrara, E., Palumbo, G., Petraglia, F., D'Agata, R., 1998. The neuroactive steroid allopregnanolone suppresses hypothalamic gonadotropin- releasing hormone release through a mechanism mediated by the gamma-aminobutyric acidA receptor. J Endocrinol 158, 121-125. Carlier, P.R., Chow, E.S., Barlow, R.L., Bloomquist, J.R., 2002. Discovery of non-zwitterionic GABA(A) receptor full agonists and a superagonist. Bioorg Med Chem Lett 12, 1985-1988. Carlson, B.X., Engblom, A.C., Kristiansen, U., Schousboe, A., Olsen, R.W., 2000. A single residue at the entrance to the first membrane-spanning domain of the gamma-aminobutyric acid type A receptor beta(2) subunit affects allosteric sensitivity to GABA and anesthetics. Mol Pharmacol 57, 474- 484. Carta, M., Mameli, M., Valenzuela, C.F., 2004. Alcohol enhances GABAergic transmission to cerebellar granule cells via an increase in Golgi cell excitability. J Neurosci 24, 3746-3751. Carter, R.B., Wood, P.L., Wieland, S., Hawkinson, J.E., Belelli, D., Lambert, J.J., White, H.S., Wolf, H.H., Mirsadeghi, S., Tahir, S.H., Bolger, M.B., Lan, N.C., Gee, K.W., 1997. Characterization of the anticonvulsant properties of ganaxolone (CCD 1042; 3alpha-hydroxy-3beta-methyl-5alpha-pregnan-20- one), a selective, high-affinity, steroid modulator of the gamma-aminobutyric acid(A) receptor. J Pharmacol Exp Ther 280, 1284-1295. Cestari, I.N., Min, K.T., Kulli, J.C., Yang, J., 2000. Identification of an amino acid defining the distinct properties of murine beta1 and beta3 subunit-containing GABA(A) receptors. J Neurochem 74, 827- 838. Chang, C.S., Olcese, R., Olsen, R.W., 2003. A single M1 residue in the beta2 subunit alters channel gating of GABAA receptor in anesthetic modulation and direct activation. J Biol Chem 278, 42821- 42828. Chebib, M., Johnston, G.A., 2000. GABA-Activated ligand gated ion channels: medicinal chemistry and molecular biology. J Med Chem 43, 1427-1447.

57 Cheney, D.L., Uzunov, D., Costa, E., Guidotti, A., 1995. Gas chromatographic-mass fragmentographic quantitation of 3 alpha-hydroxy-5 alpha-pregnan-20-one (allopregnanolone) and its precursors in blood and brain of adrenalectomized and castrated rats. J Neurosci 15, 4641-4650. Collins, I., Moyes, C., Davey, W.B., Rowley, M., Bromidge, F.A., Quirk, K., Atack, J.R., McKernan, R.M., Thompson, S.A., Wafford, K., Dawson, G.R., Pike, A., Sohal, B., Tsou, N.N., Ball, R.G., Castro, J.L., 2002. 3-Heteroaryl-2-pyridones: benzodiazepine site ligands with functional delectivity for alpha 2/alpha 3-subtypes of human GABA(A) receptor-ion channels. J Med Chem 45, 1887-1900. Collinson, N., Kuenzi, F.M., Jarolimek, W., Maubach, K.A., Cothliff, R., Sur, C., Smith, A., Otu, F.M., Howell, O., Atack, J.R., McKernan, R.M., Seabrook, G.R., Dawson, G.R., Whiting, P.J., Rosahl, T.W., 2002. Enhanced learning and memory and altered GABAergic synaptic transmission in mice lacking the alpha 5 subunit of the GABAA receptor. J Neurosci 22, 5572-5580. Compagnone, N.A., Mellon, S.H., 2000. Neurosteroids: biosynthesis and function of these novel neuromodulators. Front Neuroendocrinol 21, 1-56. Concas, A., Follesa, P., Barbaccia, M.L., Purdy, R.H., Biggio, G., 1999. Physiological modulation of GABA(A) receptor plasticity by progesterone metabolites. Eur J Pharmacol 375, 225-235. Concas, A., Mostallino, M.C., Porcu, P., Follesa, P., Barbaccia, M.L., Trabucchi, M., Purdy, R.H., Grisenti, P., Biggio, G., 1998. Role of brain allopregnanolone in the plasticity of gamma-aminobutyric acid type A receptor in rat brain during pregnancy and after delivery. Proc Natl Acad Sci U S A 95, 13284-13289. Conti, F., Stuhmer, W., 1989. Quantal charge redistributions accompanying the structural transitions of sodium channels. Eur Biophys J 17, 53-59. Cooper, E.J., Johnston, G.A., Edwards, F.A., 1999. Effects of a naturally occurring neurosteroid on GABAA IPSCs during development in rat hippocampal or cerebellar slices. J Physiol 521 Pt 2, 437- 449. Corpechot, C., Young, J., Calvel, M., Wehrey, C., Veltz, J.N., Touyer, G., Mouren, M., Prasad, V.V., Banner, C., Sjovall, J., et al., 1993. Neurosteroids: 3 alpha-hydroxy-5 alpha-pregnan-20-one and its precursors in the brain, plasma, and steroidogenic glands of male and female rats. Endocrinology 133, 1003-1009. Costa, E., Cheney, D.L., Grayson, D.R., Korneyev, A., Longone, P., Pani, L., Romeo, E., Zivkovich, E., Guidotti, A., 1994. Pharmacology of neurosteroid biosynthesis. Role of the mitochondrial DBI receptor (MDR) complex. Ann N Y Acad Sci 746, 223-242. Coulter, D.A., 2001. Epilepsy-associated plasticity in gamma-aminobutyric acid receptor expression, function, and inhibitory synaptic properties. Int Rev Neurobiol 45, 237-252. Covey, D.F., Han, M., Kumar, A.S., de La Cruz, M.A., Meadows, E.S., Hu, Y., Tonnies, A., Nathan, D., Coleman, M., Benz, A., Evers, A.S., Zorumski, C.F., Mennerick, S., 2000. Neurosteroid analogues. 8. Structure-activity studies of N-acylated 17a-aza-D-homosteroid analogues of the anesthetic steroids (3alpha, 5alpha)- and (3alpha,5beta)-3-hydroxypregnan-20-one. J Med Chem 43, 3201-3204. Crawley, J.N., Glowa, J.R., Majewska, M.D., Paul, S.M., 1986. Anxiolytic activity of an endogenous . Brain Res 398, 382-385. Crestani, F., Keist, R., Fritschy, J.M., Benke, D., Vogt, K., Prut, L., Bluthmann, H., Mohler, H., Rudolph, U., 2002. Trace fear conditioning involves hippocampal alpha5 GABA(A) receptors. Proc Natl Acad Sci U S A 99, 8980-8985. Crestani, F., Lorez, M., Baer, K., Essrich, C., Benke, D., Laurent, J.P., Belzung, C., Fritschy, J.M., Luscher, B., Mohler, H., 1999. Decreased GABAA-receptor clustering results in enhanced anxiety and a bias for threat cues. Nat Neurosci 2, 833-839. Crestani, F., Low, K., Keist, R., Mandelli, M., Mohler, H., Rudolph, U., 2001. Molecular targets for the myorelaxant action of diazepam. Mol Pharmacol 59, 442-445. Czlonkowska, A.I., Krzascik, P., Sienkiewicz-Jarosz, H., Siemiatkowski, M., Szyndler, J., Maciejak, P., Bidzinski, A., Plaznik, A., 2001. Tolerance to the anticonvulsant activity of and allopregnanolone in a model of picrotoxin seizures. Eur J Pharmacol 425, 121-127.

58 Dalziel, J.E., Cox, G.B., Gage, P.W., Birnir, B., 1999. Mutant human alpha(1)beta(1)(T262Q) GABA(A) receptors are directly activated but not modulated by pentobarbital. Eur J Pharmacol 385, 283-286. Dascal, N., 1987. The use of Xenopus oocytes for the study of ion channels. CRC Crit Rev Biochem 22, 317-387. Davies, P.A., Hanna, M.C., Hales, T.G., Kirkness, E.F., 1997. Insensitivity to anaesthetic agents conferred by a class of GABA(A) receptor subunit. Nature 385, 820-823. Deitrich, R.A., Dunwiddie, T.V., Harris, R.A., Erwin, V.G., 1989. Mechanism of action of ethanol: initial central nervous system actions. Pharmacol Rev 41, 489-537. DeLorey, T.M., Handforth, A., Anagnostaras, S.G., Homanics, G.E., Minassian, B.A., Asatourian, A., Fanselow, M.S., Delgado-Escueta, A., Ellison, G.D., Olsen, R.W., 1998. Mice lacking the beta3 subunit of the GABAA receptor have the epilepsy phenotype and many of the behavioral characteristics of Angelman syndrome. J Neurosci 18, 8505-8514. Demirgoren, S., Majewska, M.D., Spivak, C.E., London, E.D., 1991. Receptor binding and electrophysiological effects of dehydroepiandrosterone sulfate, an antagonist of the GABAA receptor. Neuroscience 45, 127-135. Devaud, L.L., Purdy, R.H., Finn, D.A., Morrow, A.L., 1996. Sensitization of gamma-aminobutyric acidA receptors to neuroactive steroids in rats during ethanol withdrawal. J Pharmacol Exp Ther 278, 510-517. Devaud, L.L., Purdy, R.H., Morrow, A.L., 1995. The neurosteroid, 3 alpha-hydroxy-5 alpha-pregnan- 20-one, protects against bicuculline-induced seizures during ethanol withdrawal in rats. Alcohol Clin Exp Res 19, 350-355. Devor, A., Fritschy, J.M., Yarom, Y., 2001. Spatial distribution and subunit composition of GABA(A) receptors in the inferior olivary nucleus. J Neurophysiol 85, 1686-1696. Dias, R., Sheppard, W.F., Fradley, R.L., Garrett, E.M., Stanley, J.L., Tye, S.J., Goodacre, S., Lincoln, R.J., Cook, S.M., Conley, R., Hallett, D., Humphries, A.C., Thompson, S.A., Wafford, K.A., Street, L.J., Castro, J.L., Whiting, P.J., Rosahl, T.W., Atack, J.R., McKernan, R.M., Dawson, G.R., Reynolds, D.S., 2005. Evidence for a significant role of alpha 3-containing GABAA receptors in mediating the anxiolytic effects of benzodiazepines. J Neurosci 25, 10682-10688. Drafts, B.C., Fisher, J.L., 2006. Identification of structures within GABAA receptor alpha subunits that regulate the agonist action of pentobarbital. J Pharmacol Exp Ther 318, 1094-1101. Draguhn, A., Verdorn, T.A., Ewert, M., Seeburg, P.H., Sakmann, B., 1990. Functional and molecular distinction between recombinant rat GABAA receptor subtypes by Zn2+. Neuron 5, 781-788. Drew, C.A., Johnston, G.A., Weatherby, R.P., 1984. Bicuculline-insensitive GABA receptors: studies on the binding of (-)- to rat cerebellar membranes. Neurosci Lett 52, 317-321. Droogleever Fortuyn, H.A., van Broekhoven, F., Span, P.N., Backstrom, T., Zitman, F.G., Verkes, R.J., 2004. Effects of PhD examination stress on allopregnanolone and cortisol plasma levels and peripheral benzodiazepine receptor density. Psychoneuroendocrinology 29, 1341-1344. Duncan, J.S., 1999. Positron emission tomography receptor studies in epilepsy. Rev Neurol (Paris) 155, 482-488. Ebert, B., i Storustovu, S., Mortensen, M., Frolund, B., 2002. Characterization of GABA(A) receptor ligands in the rat cortical wedge preparation: evidence for action at extrasynaptic receptors? Br J Pharmacol 137, 1-8. Ebert, B., Wafford, K.A., Whiting, P.J., Krogsgaard-Larsen, P., Kemp, J.A., 1994. Molecular pharmacology of gamma-aminobutyric acid type A receptor agonists and partial agonists in oocytes injected with different alpha, beta, and gamma receptor subunit combinations. Mol Pharmacol 46, 957- 963. Edgar, D.M., Seidel, W.F., Gee, K.W., Lan, N.C., Field, G., Xia, H., Hawkinson, J.E., Wieland, S., Carter, R.B., Wood, P.L., 1997. CCD-3693: an orally bioavailable analog of the endogenous neuroactive steroid, pregnanolone, demonstrates potent sedative hypnotic actions in the rat. J Pharmacol Exp Ther 282, 420-429.

59 Edwards, H.E., Vimal, S., Burnham, W.M., 2005. The acute anticonvulsant effects of deoxycorticosterone in developing rats: role of metabolites and mineralocorticoid-receptor responses. Epilepsia 46, 1888-1897. Eisenman, L.N., He, Y., Fields, C., Zorumski, C.F., Mennerick, S., 2003. Activation dependent properties of pregnenolone sulfate inhibition of GABAA receptor-mediated current. J Physiol 550, 679- 691. Erlander, M.G., Tillakaratne, N.J., Feldblum, S., Patel, N., Tobin, A.J., 1991. Two genes encode distinct glutamate decarboxylases. Neuron 7, 91-100. Esclapez, M., Tillakaratne, N.J., Kaufman, D.L., Tobin, A.J., Houser, C.R., 1994. Comparative localization of two forms of glutamic acid decarboxylase and their mRNAs in rat brain supports the concept of functional differences between the forms. J Neurosci 14, 1834-1855. Essrich, C., Lorez, M., Benson, J.A., Fritschy, J.M., Luscher, B., 1998. Postsynaptic clustering of major GABAA receptor subtypes requires the gamma 2 subunit and gephyrin. Nat Neurosci 1, 563-571. Evans, R.H., 1979. Potentiation of the effects of GABA by pentobarbitone. Brain Res 171, 113-120. Farr, S.A., Uezu, K., Creonte, T.A., Flood, J.F., Morley, J.E., 2000. Modulation of memory processing in the cingulate cortex of mice. Pharmacol Biochem Behav 65, 363-368. Farrant, M., Nusser, Z., 2005. Variations on an inhibitory theme: phasic and tonic activation of GABA(A) receptors. Nat Rev Neurosci 6, 215-229. Farrar, S.J., Whiting, P.J., Bonnert, T.P., McKernan, R.M., 1999. Stoichiometry of a ligand-gated ion channel determined by fluorescence energy transfer. J Biol Chem 274, 10100-10104. Ferrara, G., Serra, M., Zammaretti, F., Pisu, M.G., Panzica, G.C., Biggio, G., Eva, C., 2001. Increased expression of the neuropeptide Y receptor Y(1) gene in the medial amygdala of transgenic mice induced by long-term treatment with progesterone or allopregnanolone. J Neurochem 79, 417-425. Fick, D.B., Hawkinson, J.E., Acosta-Burruel, M., Nelson, T., Prins, T., Vanover, K.E., Carter, R.B., Gee, K.W., Lan, N., Hogenkamp, D.J., 1999. Sodium S-(3 α-hydroxypregnan-20-on-21-yl) thiosulfates: allosteric modulation modultators of GABAA receptor. Book of abstractss, 217th ACS national meeting, Anaheim, CA, United States March 21-25. Fish, E.W., Faccidomo, S., DeBold, J.F., Miczek, K.A., 2001. Alcohol, allopregnanolone and aggression in mice. Psychopharmacology (Berl) 153, 473-483. Fisher, J.L., Macdonald, R.L., 1997. Single channel properties of recombinant GABAA receptors containing gamma 2 or delta subtypes expressed with alpha 1 and beta 3 subtypes in mouse L929 cells. J Physiol 505 (Pt 2), 283-297. Flood, J.F., Morley, J.E., Roberts, E., 1992. Memory-enhancing effects in male mice of pregnenolone and steroids metabolically derived from it. Proc Natl Acad Sci U S A 89, 1567-1571. Flood, J.F., Morley, J.E., Roberts, E., 1995. Pregnenolone sulfate enhances post-training memory processes when injected in very low doses into limbic system structures: the amygdala is by far the most sensitive. Proc Natl Acad Sci U S A 92, 10806-10810. Flood, J.F., Smith, G.E., Roberts, E., 1988. Dehydroepiandrosterone and its sulfate enhance memory retention in mice. Brain Res 447, 269-278. Follesa, P., Concas, A., Porcu, P., Sanna, E., Serra, M., Mostallino, M.C., Purdy, R.H., Biggio, G., 2001. Role of allopregnanolone in regulation of GABA(A) receptor plasticity during long-term exposure to and withdrawal from progesterone. Brain Res Brain Res Rev 37, 81-90. Follesa, P., Floris, S., Tuligi, G., Mostallino, M.C., Concas, A., Biggio, G., 1998. Molecular and functional adaptation of the GABA(A) receptor complex during pregnancy and after delivery in the rat brain. Eur J Neurosci 10, 2905-2912. Franks, N.P., Lieb, W.R., 1994. Molecular and cellular mechanisms of general anaesthesia. Nature 367, 607-614. Frederickson, C.J., 1989. Neurobiology of zinc and zinc-containing neurons. Int Rev Neurobiol 31, 145-238. Frederickson, C.J., Bush, A.I., 2001. Synaptically released zinc: physiological functions and pathological effects. Biometals 14, 353-366.

60 Fritschy, J.M., Brunig, I., 2003. Formation and plasticity of GABAergic synapses: physiological mechanisms and pathophysiological implications. Pharmacol Ther 98, 299-323. Frye, C.A., Mermelstein, P.G., DeBold, J.F., 1992. Evidence for a non-genomic action of progestins on sexual receptivity in hamster ventral tegmental area but not hypothalamus. Brain Res 578, 87-93. Frye, C.A., Van Keuren, K.R., Erskine, M.S., 1996. Behavioral effects of 3 alpha-androstanediol. I: Modulation of sexual receptivity and promotion of GABA-stimulated chloride flux. Behav Brain Res 79, 109-118. Frye, G.D., Chapin, R.E., Vogel, R.A., Mailman, R.B., Kilts, C.D., Mueller, R.A., Breese, G.R., 1981. Effects of acute and chronic 1,3-butanediol treatment on central nervous system function: a comparison with ethanol. J Pharmacol Exp Ther 216, 306-314. Gasior, M., Carter, R.B., Witkin, J.M., 1999. Neuroactive steroids: potential therapeutic use in neurological and psychiatric disorders. Trends Pharmacol Sci 20, 107-112. Gee, K.W., Lan, N.C., 1991. Gamma-aminobutyric acidA receptor complexes in rat frontal cortex and spinal cord show differential responses to steroid modulation. Mol Pharmacol 40, 995-999. Gibbs, J.W., 3rd, Sombati, S., DeLorenzo, R.J., Coulter, D.A., 1997. Physiological and pharmacological alterations in postsynaptic GABA(A) receptor function in a hippocampal culture model of chronic spontaneous seizures. J Neurophysiol 77, 2139-2152. Gibbs, T.T., Russek, S.J., Farb, D.H., 2006. Sulfated steroids as endogenous neuromodulators. Pharmacol Biochem Behav 84, 555-567. Givalois, L., Grinevich, V., Li, S., Garcia-De-Yebenes, E., Pelletier, G., 1998. The octadecaneuropeptide-induced response of corticotropin-releasing hormone messenger RNA levels is mediated by GABA(A) receptors and modulated by endogenous steroids. Neuroscience 85, 557-567. Goodchild, C.S., Robinson, A., Nadeson, R., 2001. Antinociceptive properties of neurosteroids IV: pilot study demonstrating the analgesic effects of alphadolone administered orally to humans. Br J Anaesth 86, 528-534. Grishkovskaya, I., Avvakumov, G.V., Sklenar, G., Dales, D., Hammond, G.L., Muller, Y.A., 2000. Crystal structure of human sex hormone-binding globulin: steroid transport by a laminin G-like domain. Embo J 19, 504-512. Grobin, A.C., Morrow, A.L., 2000. 3alpha-hydroxy-5alpha-pregnan-20-one exposure reduces GABA(A) receptor alpha4 subunit mRNA levels. Eur J Pharmacol 409, R1-2. Grunewald, R.A., Aliberti, V., Panayiotopoulos, C.P., 1992. Exacerbation of typical absence seizures by progesterone. Seizure 1, 137-138. Gulinello, M., Gong, Q.H., Li, X., Smith, S.S., 2001. Short-term exposure to a neuroactive steroid increases alpha4 GABA(A) receptor subunit levels in association with increased anxiety in the female rat. Brain Res 910, 55-66. Gyermek, L., Iriarte, J., Crabbe, P., 1968. Steroids. CCCX. Structure-activity relationship of some steroidal hypnotic agents. J Med Chem 11, 117-125. Gyermek, L., Soyka, L.F., 1975. Steroid anesthetics. Anesthesiology 42, 331-344. Haage, D., Backstrom, T., Johansson, S., 2005. Interaction between allopregnanolone and pregnenolone sulfate in modulating GABA-mediated synaptic currents in neurons from the rat medial preoptic nucleus. Brain Res 1033, 58-67. Haage, D., Johansson, S., 1999. Neurosteroid modulation of synaptic and GABA-evoked currents in neurons from the rat medial preoptic nucleus. J Neurophysiol 82, 143-151. Hadingham, K.L., Wafford, K.A., Thompson, S.A., Palmer, K.J., Whiting, P.J., 1995. Expression and pharmacology of human GABAA receptors containing gamma 3 subunits. Eur J Pharmacol 291, 301- 309. Hadingham, K.L., Wingrove, P.B., Wafford, K.A., Bain, C., Kemp, J.A., Palmer, K.J., Wilson, A.W., Wilcox, A.S., Sikela, J.M., Ragan, C.I., et al., 1993. Role of the beta subunit in determining the pharmacology of human gamma-aminobutyric acid type A receptors. Mol Pharmacol 44, 1211-1218. Halbreich, U., 2003. The etiology, biology, and evolving pathology of premenstrual syndromes. Psychoneuroendocrinology 28 Suppl 3, 55-99.

61 Hamann, M., Rossi, D.J., Attwell, D., 2002. Tonic and spillover inhibition of granule cells control information flow through cerebellar cortex. Neuron 33, 625-633. Hamilton, N.M., 2002. Interaction of steroids with the GABA(A) receptor. Curr Top Med Chem 2, 887- 902. Han, M., Zorumski, C.F., Covey, D.F., 1996. Neurosteroid analogues. 4. The effect of methyl substitution at the C-5 and C-10 positions of neurosteroids on electrophysiological activity at GABAA receptors. J Med Chem 39, 4218-4232. Harney, S.C., Frenguelli, B.G., Lambert, J.J., 2003. Phosphorylation influences neurosteroid modulation of synaptic GABAA receptors in rat CA1 and dentate gyrus neurones. Neuropharmacology 45, 873-883. Harris, R.A., Proctor, W.R., McQuilkin, S.J., Klein, R.L., Mascia, M.P., Whatley, V., Whiting, P.J., Dunwiddie, T.V., 1995. Ethanol increases GABAA responses in cells stably transfected with receptor subunits. Alcohol Clin Exp Res 19, 226-232. Harris, R.A., Valenzuela, C.F., Brozowski, S., Chuang, L., Hadingham, K., Whiting, P.J., 1998. Adaptation of gamma-aminobutyric acid type A receptors to alcohol exposure: studies with stably transfected cells. J Pharmacol Exp Ther 284, 180-188. Havlikova, H., Hill, M., Kancheva, L., Vrbikova, J., Pouzar, V., Cerny, I., Kancheva, R., Starka, L., 2006. Serum profiles of free and conjugated neuroactive pregnanolone isomers in nonpregnant women of fertile age. J Clin Endocrinol Metab 91, 3092-3099. Hawkinson, J.E., Kimbrough, C.L., Belelli, D., Lambert, J.J., Purdy, R.H., Lan, N.C., 1994. Correlation of neuroactive steroid modulation of [35S]t-butylbicyclophosphorothionate and [3H]flunitrazepam binding and gamma-aminobutyric acidA receptor function. Mol Pharmacol 46, 977-985. Hawkinson, J.E., Kimbrough, C.L., McCauley, L.D., Bolger, M.B., Lan, N.C., Gee, K.W., 1994. The neuroactive steroid 3 alpha-hydroxy-5 beta-pregnan-20-one is a two-component modulator of ligand binding to the GABAA receptor. Eur J Pharmacol 269, 157-163. Herbison, A.E., 2001. Physiological roles for the neurosteroid allopregnanolone in the modulation of brain function during pregnancy and parturition. Prog Brain Res 133, 39-47. Hill-Venning, C., Belelli, D., Peters, J.A., Lambert, J.J., 1997. Subunit-dependent interaction of the etomidate with the gamma-aminobutyric acid type A receptor. Br J Pharmacol 120, 749-756. Hill-Venning, C., Peters, J.A., Callachan, H., Lambert, J.J., Gemmell, D.K., Anderson, A., Byford, A., Hamilton, N., Hill, D.R., Marshall, R.J., Campbell, A.C., 1996. The anaesthetic action and modulation of GABAA receptor activity by the novel water-soluble aminosteroid Org 20599. Neuropharmacology 35, 1209-1222. Hill, D.R., Bowery, N.G., 1981. 3H-baclofen and 3H-GABA bind to bicuculline-insensitive GABA B sites in rat brain. Nature 290, 149-152. Hill, M., Bicikova, M., Parizek, A., Havlikova, H., Klak, J., Fajt, T., Meloun, M., Cibula, D., Cegan, A., Sulcova, J., Hampl, R., Starka, L., 2001. Neuroactive steroids, their precursors and polar conjugates during parturition and postpartum in maternal blood: 2. Time profiles of pregnanolone isomers. J Steroid Biochem Mol Biol 78, 51-57. Hillen, T., Lun, A., Reischies, F.M., Borchelt, M., Steinhagen-Thiessen, E., Schaub, R.T., 2000. DHEA-S plasma levels and incidence of Alzheimer's disease. Biol Psychiatry 47, 161-163. Hogenkamp, D.J., Tahir, S.H., Hawkinson, J.E., Upasani, R.B., Alauddin, M., Kimbrough, C.L., Acosta-Burruel, M., Whittemore, E.R., Woodward, R.M., Lan, N.C., Gee, K.W., Bolger, M.B., 1997. Synthesis and in vitro activity of 3 beta-substituted-3 alpha-hydroxypregnan-20-ones: allosteric modulators of the GABAA receptor. J Med Chem 40, 61-72. Hosie, A.M., Dunne, E.L., Harvey, R.J., Smart, T.G., 2003. Zinc-mediated inhibition of GABA(A) receptors: discrete binding sites underlie subtype specificity. Nat Neurosci 6, 362-369. Hosie, A.M., Wilkins, M.E., da Silva, H.M., Smart, T.G., 2006. Endogenous neurosteroids regulate GABAA receptors through two discrete transmembrane sites. Nature 444, 486-489. Hutcheon, B., Morley, P., Poulter, M.O., 2000. Developmental change in GABAA receptor desensitization kinetics and its role in synapse function in rat cortical neurons. J Physiol 522 Pt 1, 3-17.

62 Imamura, M., Prasad, C., 1998. Modulation of GABA-gated chloride ion influx in the brain by dehydroepiandrosterone and its metabolites. Biochem Biophys Res Commun 243, 771-775. Johansson, I.M., Birzniece, V., Lindblad, C., Olsson, T., Backstrom, T., 2002. Allopregnanolone inhibits learning in the Morris water maze. Brain Res 934, 125-131. Jones, M.V., Westbrook, G.L., 1995. Desensitized states prolong GABAA channel responses to brief agonist pulses. Neuron 15, 181-191. Kancheva, L., Havlikova, H., Hill, M., Vrbikova, J., Starka, L., 2007. Neuroactive steroids in adult men. 4th International Meeting- Steroids and Nervous system, Torino, Italy, Abstract book, p152. Kellenberger, S., Malherbe, P., Sigel, E., 1992. Function of the alpha 1 beta 2 gamma 2S gamma- aminobutyric acid type A receptor is modulated by protein kinase C via multiple phosphorylation sites. J Biol Chem 267, 25660-25663. Kennedy, R.T., Thompson, J.E., Vickroy, T.W., 2002. In vivo monitoring of amino acids by direct sampling of brain extracellular fluid at ultralow flow rates and capillary electrophoresis. J Neurosci Methods 114, 39-49. Kerr, D.I., Ong, J., 1992. GABA agonists and antagonists. Med Res Rev 12, 593-636. Khisti, R.T., Mandhane, S.N., Chopde, C.T., 1998. The neurosteroid 3 alpha-hydroxy-5 alpha-pregnan- 20-one induces catalepsy in mice. Neurosci Lett 251, 85-88. Kittler, J.T., Wang, J., Connolly, C.N., Vicini, S., Smart, T.G., Moss, S.J., 2000. Analysis of GABAA receptor assembly in mammalian cell lines and hippocampal neurons using gamma 2 subunit green fluorescent protein chimeras. Mol Cell Neurosci 16, 440-452. Klangkalya, B., Chan, A., 1988. Structure-activity relationships of steroid hormones on muscarinic receptor binding. J Steroid Biochem 29, 111-118. Klausberger, T., Ehya, N., Fuchs, K., Fuchs, T., Ebert, V., Sarto, I., Sieghart, W., 2001. Detection and binding properties of GABA(A) receptor assembly intermediates. J Biol Chem 276, 16024-16032. Knight, A.R., Stephenson, F.A., Tallman, J.F., Ramabahdran, T.V., 2000. Monospecific antibodies as probes for the stoichiometry of recombinant GABA(A) receptors. Receptors Channels 7, 213-226. Kokate, T.G., Svensson, B.E., Rogawski, M.A., 1994. Anticonvulsant activity of neurosteroids: correlation with gamma-aminobutyric acid-evoked chloride current potentiation. J Pharmacol Exp Ther 270, 1223-1229. Koksma, J.J., van Kesteren, R.E., Rosahl, T.W., Zwart, R., Smit, A.B., Luddens, H., Brussaard, A.B., 2003. Oxytocin regulates neurosteroid modulation of GABA(A) receptors in supraoptic nucleus around parturition. J Neurosci 23, 788-797. Krishek, B.J., Moss, S.J., Smart, T.G., 1996. A functional comparison of the antagonists bicuculline and picrotoxin at recombinant GABAA receptors. Neuropharmacology 35, 1289-1298. Krishek, B.J., Moss, S.J., Smart, T.G., 1998. Interaction of H+ and Zn2+ on recombinant and native rat neuronal GABAA receptors. J Physiol 507 (Pt 3), 639-652. Krogsgaard-Larsen, P., Johnston, G.A., Lodge, D., Curtis, D.R., 1977. A new class of GABA agonist. Nature 268, 53-55. Kunig, G., Leenders, K.L., Sanchez-Pernaute, R., Antonini, A., Vontobel, P., Verhagen, A., Gunther, I., 2000. Benzodiazepine receptor binding in Huntington's disease: [11C] uptake measured using positron emission tomography. Ann Neurol 47, 644-648. Lambert, J.J., Belelli, D., Harney, S.C., Peters, J.A., Frenguelli, B.G., 2001. Modulation of native and recombinant GABA(A) receptors by endogenous and synthetic neuroactive steroids. Brain Res Brain Res Rev 37, 68-80. Lambert, J.J., Belelli, D., Hill-Venning, C., Peters, J.A., 1995. Neurosteroids and GABAA receptor function. Trends Pharmacol Sci 16, 295-303. Lambert, J.J., Harney, S.C., Belelli, D., Peters, J.A., 2001. Neurosteroid modulation of recombinant and synaptic GABAA receptors. Int Rev Neurobiol 46, 177-205. Lanthier, A., Patwardhan, V.V., 1986. Sex steroids and 5-en-3 beta-hydroxysteroids in specific regions of the human brain and cranial nerves. J Steroid Biochem 25, 445-449. Laubach, G.D., P'An, S.Y., Rudel, H.W., 1955. Steroid anesthetic agent. Science 122, 78.

63 Lerma, J., Herranz, A.S., Herreras, O., Abraira, V., Martin del Rio, R., 1986. In vivo determination of extracellular concentration of amino acids in the rat hippocampus. A method based on brain dialysis and computerized analysis. Brain Res 384, 145-155. Leroy, C., Poisbeau, P., Keller, A.F., Nehlig, A., 2004. Pharmacological plasticity of GABA(A) receptors at dentate gyrus synapses in a rat model of temporal lobe epilepsy. J Physiol 557, 473-487. Lewis, D.A., 2000. GABAergic local circuit neurons and prefrontal cortical dysfunction in schizophrenia. Brain Res Brain Res Rev 31, 270-276. Lin, Y.F., Angelotti, T.P., Dudek, E.M., Browning, M.D., Macdonald, R.L., 1996. Enhancement of recombinant alpha 1 beta 1 gamma 2L gamma-aminobutyric acidA receptor whole-cell currents by protein kinase C is mediated through phosphorylation of both beta 1 and gamma 2L subunits. Mol Pharmacol 50, 185-195. Lofgren, M., Johansson, I.M., Meyerson, B., Lundgren, P., Backstrom, T., 2006. Progesterone withdrawal effects in the open field test can be predicted by elevated plus maze performance. Horm Behav 50, 208-215. Low, K., Crestani, F., Keist, R., Benke, D., Brunig, I., Benson, J.A., Fritschy, J.M., Rulicke, T., Bluethmann, H., Mohler, H., Rudolph, U., 2000. Molecular and neuronal substrate for the selective attenuation of anxiety. Science 290, 131-134. Luddens, H., Pritchett, D.B., Kohler, M., Killisch, I., Keinanen, K., Monyer, H., Sprengel, R., Seeburg, P.H., 1990. Cerebellar GABAA receptor selective for a behavioural alcohol antagonist. Nature 346, 648-651. Macdonald, R.L., Olsen, R.W., 1994. GABAA receptor channels. Annu Rev Neurosci 17, 569-602. Maione, S., Berrino, L., Vitagliano, S., Leyva, J., Rossi, F., 1992. Pregnenolone sulfate increases the convulsant potency of N-methyl-D-aspartate in mice. Eur J Pharmacol 219, 477-479. Maitra, R., Reynolds, J.N., 1998. Modulation of GABA(A) receptor function by neuroactive steroids: evidence for heterogeneity of steroid sensitivity of recombinant GABA(A) receptor isoforms. Can J Physiol Pharmacol 76, 909-920. Maitra, R., Reynolds, J.N., 1999. Subunit dependent modulation of GABAA receptor function by neuroactive steroids. Brain Res 819, 75-82. Majewska, M.D., 1987. Antagonist-type interaction of with the GABA receptor- coupled chloride channel. Brain Res 418, 377-382. Majewska, M.D., 1992. Neurosteroids: endogenous bimodal modulators of the GABAA receptor. Mechanism of action and physiological significance. Prog Neurobiol 38, 379-395. Majewska, M.D., Demirgoren, S., Spivak, C.E., London, E.D., 1990. The neurosteroid dehydroepiandrosterone sulfate is an allosteric antagonist of the GABAA receptor. Brain Res 526, 143- 146. Majewska, M.D., Harrison, N.L., Schwartz, R.D., Barker, J.L., Paul, S.M., 1986. Steroid hormone metabolites are barbiturate-like modulators of the GABA receptor. Science 232, 1004-1007. Majewska, M.D., Mienville, J.M., Vicini, S., 1988. Neurosteroid pregnenolone sulfate antagonizes electrophysiological responses to GABA in neurons. Neurosci Lett 90, 279-284. Malizia, A.L., 1999. What do brain imaging studies tell us about anxiety disorders? J Psychopharmacol 13, 372-378. Marshall, F.H., Stratton, S.C., Mullings, J., Ford, E., Worton, S.P., Oakley, N.R., Hagan, R.M., 1997. Development of tolerance in mice to the sedative effects of the neuroactive steroid following chronic exposure. Pharmacol Biochem Behav 58, 1-8. Mayo, W., Dellu, F., Robel, P., Cherkaoui, J., Le Moal, M., Baulieu, E.E., Simon, H., 1993. Infusion of neurosteroids into the nucleus basalis magnocellularis affects cognitive processes in the rat. Brain Res 607, 324-328. McCauley, L.D., Liu, V., Chen, J.S., Hawkinson, J.E., Lan, N.C., Gee, K.W., 1995. Selective actions of certain neuroactive pregnanediols at the gamma-aminobutyric acid type A receptor complex in rat brain. Mol Pharmacol 47, 354-362. McEwen, B.S., 1991. Non-genomic and genomic effects of steroids on neural activity. Trends Pharmacol Sci 12, 141-147.

64 McKernan, R.M., Rosahl, T.W., Reynolds, D.S., Sur, C., Wafford, K.A., Atack, J.R., Farrar, S., Myers, J., Cook, G., Ferris, P., Garrett, L., Bristow, L., Marshall, G., Macaulay, A., Brown, N., Howell, O., Moore, K.W., Carling, R.W., Street, L.J., Castro, J.L., Ragan, C.I., Dawson, G.R., Whiting, P.J., 2000. Sedative but not anxiolytic properties of benzodiazepines are mediated by the GABA(A) receptor alpha1 subtype. Nat Neurosci 3, 587-592. McKernan, R.M., Wafford, K., Quirk, K., Hadingham, K.L., Harley, E.A., Ragan, C.I., Whiting, P.J., 1995. The pharmacology of the benzodiazepine site of the GABA-A receptor is dependent on the type of gamma-subunit present. J Recept Signal Transduct Res 15, 173-183. McKernan, R.M., Whiting, P.J., 1996. Which GABAA-receptor subtypes really occur in the brain? Trends Neurosci 19, 139-143. Melchior, C.L., Ritzmann, R.F., 1996. Neurosteroids block the memory-impairing effects of ethanol in mice. Pharmacol Biochem Behav 53, 51-56. Mellon, S.H., Griffin, L.D., Compagnone, N.A., 2001. Biosynthesis and action of neurosteroids. Brain Res Brain Res Rev 37, 3-12. Mennerick, S., He, Y., Jiang, X., Manion, B.D., Wang, M., Shute, A., Benz, A., Evers, A.S., Covey, D.F., Zorumski, C.F., 2004. Selective antagonism of 5alpha-reduced neurosteroid effects at GABA(A) receptors. Mol Pharmacol 65, 1191-1197. Mensah-Nyagan, A.G., Do-Rego, J.L., Beaujean, D., Luu-The, V., Pelletier, G., Vaudry, H., 1999. Neurosteroids: expression of steroidogenic enzymes and regulation of steroid biosynthesis in the central nervous system. Pharmacol Rev 51, 63-81. Methfessel, C., Witzemann, V., Takahashi, T., Mishina, M., Numa, S., Sakmann, B., 1986. Patch clamp measurements on Xenopus laevis oocytes: currents through endogenous channels and implanted acetylcholine receptor and sodium channels. Pflugers Arch 407, 577-588. Miczek, K.A., Fish, E.W., De Bold, J.F., 2003. Neurosteroids, GABAA receptors, and escalated aggressive behavior. Horm Behav 44, 242-257. Mienville, J.M., Vicini, S., 1989. Pregnenolone sulfate antagonizes GABAA receptor-mediated currents via a reduction of channel opening frequency. Brain Res 489, 190-194. Mihalek, R.M., Banerjee, P.K., Korpi, E.R., Quinlan, J.J., Firestone, L.L., Mi, Z.P., Lagenaur, C., Tretter, V., Sieghart, W., Anagnostaras, S.G., Sage, J.R., Fanselow, M.S., Guidotti, A., Spigelman, I., Li, Z., DeLorey, T.M., Olsen, R.W., Homanics, G.E., 1999. Attenuated sensitivity to neuroactive steroids in gamma-aminobutyrate type A receptor delta subunit knockout mice. Proc Natl Acad Sci U S A 96, 12905-12910. Mitchell, S.J., Silver, R.A., 2003. Shunting inhibition modulates neuronal gain during synaptic excitation. Neuron 38, 433-445. Mody, I., De Koninck, Y., Otis, T.S., Soltesz, I., 1994. Bridging the cleft at GABA synapses in the brain. Trends Neurosci 17, 517-525. Mohler, H., Crestani, F., Rudolph, U., 2001. GABA(A)-receptor subtypes: a new pharmacology. Curr Opin Pharmacol 1, 22-25. Mohler, H., Fritschy, J.M., Crestani, F., Hensch, T., Rudolph, U., 2004. Specific GABA(A) circuits in brain development and therapy. Biochem Pharmacol 68, 1685-1690. Mohler, H., Fritschy, J.M., Luscher, B., Rudolph, U., Benson, J., Benke, D., 1996. The GABAA receptors. From subunits to diverse functions. Ion Channels 4, 89-113. Mohler, H., Fritschy, J.M., Rudolph, U., 2002. A new benzodiazepine pharmacology. J Pharmacol Exp Ther 300, 2-8. Molnar, P., Nadler, J.V., 2001. Lack of effect of mossy fiber-released zinc on granule cell GABA(A) receptors in the pilocarpine model of epilepsy. J Neurophysiol 85, 1932-1940. Morrow, A.L., VanDoren, M.J., Penland, S.N., Matthews, D.B., 2001. The role of GABAergic neuroactive steroids in ethanol action, tolerance and dependence. Brain Res Brain Res Rev 37, 98-109. Moss, S.J., Smart, T.G., Blackstone, C.D., Huganir, R.L., 1992. Functional modulation of GABAA receptors by cAMP-dependent protein phosphorylation. Science 257, 661-665.

65 Mozrzymas, J.W., Zarnowska, E.D., Pytel, M., Mercik, K., 2003. Modulation of GABA(A) receptors by hydrogen ions reveals synaptic GABA transient and a crucial role of the desensitization process. J Neurosci 23, 7981-7992. Nadeson, R., Goodchild, C.S., 2001. Antinociceptive properties of neurosteroids III: experiments with alphadolone given intravenously, intraperitoneally, and intragastrically. Br J Anaesth 86, 704-708. Nakashima, Y.M., Todorovic, S.M., Covey, D.F., Lingle, C.J., 1998. The anesthetic steroid (+)-3alpha- hydroxy-5alpha-androstane-17beta-carbonitrile blocks N-, Q-, and R-type, but not L- and P-type, high voltage-activated Ca2+ current in hippocampal and dorsal root ganglion neurons of the rat. Mol Pharmacol 54, 559-568. Nasman, B., Olsson, T., Backstrom, T., Eriksson, S., Grankvist, K., Viitanen, M., Bucht, G., 1991. Serum dehydroepiandrosterone sulfate in Alzheimer's disease and in multi-infarct dementia. Biol Psychiatry 30, 684-690. Nayeem, N., Green, T.P., Martin, I.L., Barnard, E.A., 1994. Quaternary structure of the native GABAA receptor determined by electron microscopic image analysis. J Neurochem 62, 815-818. Nguyen, Q., Sapp, D.W., Van Ness, P.C., Olsen, R.W., 1995. Modulation of GABAA receptor binding in human brain by neuroactive steroids: species and brain regional differences. Synapse 19, 77-87. Nohria, V., Giller, E., 2007. Ganaxolone. Neurother 4, 102-105. Nusser, Z., Cull-Candy, S., Farrant, M., 1997. Differences in synaptic GABA(A) receptor number underlie variation in GABA mini amplitude. Neuron 19, 697-709. Nusser, Z., Mody, I., 2002. Selective modulation of tonic and phasic inhibitions in dentate gyrus granule cells. J Neurophysiol 87, 2624-2628. Nusser, Z., Sieghart, W., Somogyi, P., 1998. Segregation of different GABAA receptors to synaptic and extrasynaptic membranes of cerebellar granule cells. J Neurosci 18, 1693-1703. Nutt, D.J., Malizia, A.L., 2001. New insights into the role of the GABA(A)-benzodiazepine receptor in psychiatric disorder. Br J Psychiatry 179, 390-396. Olney, J.W., Labruyere, J., Wang, G., Wozniak, D.F., Price, M.T., Sesma, M.A., 1991. NMDA antagonist : mechanism and prevention. Science 254, 1515-1518. Olsen, R.W., DeLorey, T.M., Gordey, M., Kang, M.H., 1999. GABA receptor function and epilepsy. Adv Neurol 79, 499-510. Park-Chung, M., Malayev, A., Purdy, R.H., Gibbs, T.T., Farb, D.H., 1999. Sulfated and unsulfated steroids modulate gamma-aminobutyric acidA receptor function through distinct sites. Brain Res 830, 72-87. Parry, B.L., 2001. The role of central serotonergic dysfunction in the aetiology of premenstrual dysphoric disorder: therapeutic implications. CNS Drugs 15, 277-285. Patte-Mensah, C., Kappes, V., Freund-Mercier, M.J., Tsutsui, K., Mensah-Nyagan, A.G., 2003. Cellular distribution and bioactivity of the key steroidogenic enzyme, cytochrome P450side chain cleavage, in sensory neural pathways. J Neurochem 86, 1233-1246. Paul, S.M., Purdy, R.H., 1992. Neuroactive steroids. Faseb J 6, 2311-2322. Penning, T.M., Sharp, R.B., Krieger, N.R., 1985. Purification and properties of 3 alpha-hydroxysteroid dehydrogenase from rat brain cytosol. Inhibition by nonsteroidal anti-inflammatory drugs and progestins. J Biol Chem 260, 15266-15272. Pistis, M., Belelli, D., McGurk, K., Peters, J.A., Lambert, J.J., 1999. Complementary regulation of anaesthetic activation of human (alpha6beta3gamma2L) and Drosophila (RDL) GABA receptors by a single amino acid residue. J Physiol 515 (Pt 1), 3-18. Polenzani, L., Woodward, R.M., Miledi, R., 1991. Expression of mammalian gamma-aminobutyric acid receptors with distinct pharmacology in Xenopus oocytes. Proc Natl Acad Sci U S A 88, 4318-4322. Porcello, D.M., Huntsman, M.M., Mihalek, R.M., Homanics, G.E., Huguenard, J.R., 2003. Intact synaptic GABAergic inhibition and altered neurosteroid modulation of thalamic relay neurons in mice lacking delta subunit. J Neurophysiol 89, 1378-1386. Prince, R.J., Simmonds, M.A., 1992. 5 beta-pregnan-3 beta-ol-20-one, a specific antagonist at the neurosteroid site of the GABAA receptor-complex. Neurosci Lett 135, 273-275.

66 Prince, R.J., Simmonds, M.A., 1993. Differential antagonism by epipregnanolone of alphaxalone and pregnanolone potentiation of [3H]flunitrazepam binding suggests more than one class of binding site for steroids at GABAA receptors. Neuropharmacology 32, 59-63. Pritchett, D.B., Luddens, H., Seeburg, P.H., 1989. Type I and type II GABAA-benzodiazepine receptors produced in transfected cells. Science 245, 1389-1392. Puia, G., Santi, M.R., Vicini, S., Pritchett, D.B., Purdy, R.H., Paul, S.M., Seeburg, P.H., Costa, E., 1990. Neurosteroids act on recombinant human GABAA receptors. Neuron 4, 759-765. Purdy, R.H., Moore, P.H., Jr., Morrow, A.L., Paul, S.M., 1992. Neurosteroids and GABAA receptor function. Adv Biochem Psychopharmacol 47, 87-92. Purdy, R.H., Morrow, A.L., Blinn, J.R., Paul, S.M., 1990. Synthesis, metabolism, and pharmacological activity of 3 alpha-hydroxy steroids which potentiate GABA-receptor-mediated chloride ion uptake in rat cerebral cortical synaptoneurosomes. J Med Chem 33, 1572-1581. Purdy, R.H., Morrow, A.L., Moore, P.H., Jr., Paul, S.M., 1991. Stress-induced elevations of gamma- aminobutyric acid type A receptor-active steroids in the rat brain. Proc Natl Acad Sci U S A 88, 4553- 4557. Purvez, D., Augustine, G.J., Fitzpatrick, D., Hall, W.C., LaMantia, A.S., McNamara, J.O., Williams, S.M., 2004. Neuroscinece. Sinauer Associates Inc.USA third edition, p143. Ragagnin, G., Rahman, M., Zingmark, E., Stromberg, J., Lundgren, P., Wang, M., Backstrom, T., 2007. Structure actvity relationship of GABAA-steroids antagonists. 4th International Meeting- Steroids and Nervous system, Torino, Italy, Abstract book, p 197. Reddy, D.S., 2003. Is there a physiological role for the neurosteroid THDOC in stress-sensitive conditions? Trends Pharmacol Sci 24, 103-106. Reddy, D.S., 2003. Pharmacology of endogenous neuroactive steroids. Crit Rev Neurobiol 15, 197-234. Reddy, D.S., Castaneda, D.C., O'Malley, B.W., Rogawski, M.A., 2004. Anticonvulsant activity of progesterone and neurosteroids in knockout mice. J Pharmacol Exp Ther 310, 230-239. Reddy, D.S., Kim, H.Y., Rogawski, M.A., 2001. Neurosteroid withdrawal model of perimenstrual catamenial epilepsy. Epilepsia 42, 328-336. Reddy, D.S., Kulkarni, S.K., 2000. Development of neurosteroid-based novel psychotropic drugs. Prog Med Chem 37, 135-175. Reddy, D.S., Rogawski, M.A., 2002. Stress-induced deoxycorticosterone-derived neurosteroids modulate GABA(A) receptor function and seizure susceptibility. J Neurosci 22, 3795-3805. Reilly, M.T., Crabbe, J.C., Rustay, N.R., Finn, D.A., 2000. Acute neuroactive steroid withdrawal in withdrawal seizure-prone and withdrawal seizure-resistant mice. Pharmacol Biochem Behav 67, 709- 717. Robertson, B., 1989. Actions of anaesthetics and on GABAA chloride channels in mammalian dorsal root ganglion neurones. Br J Pharmacol 98, 167-176. Rodgers-Neame, N.T., Covey, D.F., Hu, Y., Isenberg, K.E., Zorumski, C.F., 1992. Effects of a benz[e]indene on gamma-aminobutyric acid-gated chloride currents in cultured postnatal rat hippocampal neurons. Mol Pharmacol 42, 952-957. Rodgers, R.J., Johnson, N.J., 1998. Behaviorally selective effects of neuroactive steroids on plus-maze anxiety in mice. Pharmacol Biochem Behav 59, 221-232. Rowlett, J.K., Winger, G., Carter, R.B., Wood, P.L., Woods, J.H., Woolverton, W.L., 1999. Reinforcing and discriminative stimulus effects of the neuroactive steroids pregnanolone and Co 8-7071 in rhesus monkeys. Psychopharmacology (Berl) 145, 205-212. Rudolph, U., Crestani, F., Benke, D., Brunig, I., Benson, J.A., Fritschy, J.M., Martin, J.R., Bluethmann, H., Mohler, H., 1999. Benzodiazepine actions mediated by specific gamma-aminobutyric acid(A) receptor subtypes. Nature 401, 796-800. Rupprecht, R., 2003. Neuroactive steroids: mechanisms of action and neuropsychopharmacological properties. Psychoneuroendocrinology 28, 139-168.

67 Rupprecht, R., Berning, B., Hauser, C.A., Holsboer, F., Reul, J.M., 1996. Steroid receptor-mediated effects of neuroactive steroids: characterization of structure-activity relationship. Eur J Pharmacol 303, 227-234. Rupprecht, R., di Michele, F., Hermann, B., Strohle, A., Lancel, M., Romeo, E., Holsboer, F., 2001. Neuroactive steroids: molecular mechanisms of action and implications for neuropsychopharmacology. Brain Res Brain Res Rev 37, 59-67. Rupprecht, R., Reul, J.M., Trapp, T., van Steensel, B., Wetzel, C., Damm, K., Zieglgansberger, W., Holsboer, F., 1993. Progesterone receptor-mediated effects of neuroactive steroids. Neuron 11, 523- 530. Russell, D.W., Wilson, J.D., 1994. Steroid 5 alpha-reductase: two genes/two enzymes. Annu Rev Biochem 63, 25-61. Sanna, E., Murgia, A., Casula, A., Biggio, G., 1997. Differential subunit dependence of the actions of the general anesthetics alphaxalone and etomidate at gamma-aminobutyric acid type A receptors expressed in Xenopus laevis oocytes. Mol Pharmacol 51, 484-490. Saxena, N.C., Macdonald, R.L., 1996. Properties of putative cerebellar gamma-aminobutyric acid A receptor isoforms. Mol Pharmacol 49, 567-579. Schofield, P.R., Darlison, M.G., Fujita, N., Burt, D.R., Stephenson, F.A., Rodriguez, H., Rhee, L.M., Ramachandran, J., Reale, V., Glencorse, T.A., et al., 1987. Sequence and functional expression of the GABA A receptor shows a ligand-gated receptor super-family. Nature 328, 221-227. Schwartz, R.D., Paul, S.M., Majewska, M.D., 1986. Factors modulating the sensitivity of the GABA receptor-gated chloride ion channel. Clin Neuropharmacol 9 Suppl 4, 389-391. Schweizer, C., Balsiger, S., Bluethmann, H., Mansuy, I.M., Fritschy, J.M., Mohler, H., Luscher, B., 2003. The gamma 2 subunit of GABA(A) receptors is required for maintenance of receptors at mature synapses. Mol Cell Neurosci 24, 442-450. Sear, J.W., 1997. ORG 21465, a new water-soluble steroid hypnotic: more of the same or something different? Br J Anaesth 79, 417-419. Selye, H., 1942. Correlations between the chemical structure and the pharmacological actions of the steroids. Endocrinology 30, 437-453. Semyanov, A., Walker, M.C., Kullmann, D.M., 2003. GABA uptake regulates cortical excitability via cell type-specific tonic inhibition. Nat Neurosci 6, 484-490. Semyanov, A., Walker, M.C., Kullmann, D.M., Silver, R.A., 2004. Tonically active GABA A receptors: modulating gain and maintaining the tone. Trends Neurosci 27, 262-269. Serafini, R., Bracamontes, J., Steinbach, J.H., 2000. Structural domains of the human GABAA receptor 3 subunit involved in the actions of pentobarbital. J Physiol 524 Pt 3, 649-676. Serra, M., Pisu, M.G., Littera, M., Papi, G., Sanna, E., Tuveri, F., Usala, L., Purdy, R.H., Biggio, G., 2000. Social isolation-induced decreases in both the abundance of neuroactive steroids and GABA(A) receptor function in rat brain. J Neurochem 75, 732-740. Shen, W., Mennerick, S., Covey, D.F., Zorumski, C.F., 2000. Pregnenolone sulfate modulates inhibitory synaptic transmission by enhancing GABA(A) receptor desensitization. J Neurosci 20, 3571- 3579. Shen, W., Mennerick, S., Zorumski, E.C., Covey, D.F., Zorumski, C.F., 1999. Pregnenolone sulfate and dehydroepiandrosterone sulfate inhibit GABA-gated chloride currents in Xenopus oocytes expressing picrotoxin-insensitive GABA(A) receptors. Neuropharmacology 38, 267-271. Shu, H.J., Eisenman, L.N., Jinadasa, D., Covey, D.F., Zorumski, C.F., Mennerick, S., 2004. Slow actions of neuroactive steroids at GABAA receptors. J Neurosci 24, 6667-6675. Sieghart, W., 1992. GABAA receptors: ligand-gated Cl- ion channels modulated by multiple drug- binding sites. Trends Pharmacol Sci 13, 446-450. Sieghart, W., 1995. Structure and pharmacology of gamma-aminobutyric acidA receptor subtypes. Pharmacol Rev 47, 181-234. Sieghart, W., 2000. Unraveling the function of GABA(A) receptor subtypes. Trends Pharmacol Sci 21, 411-413.

68 Sigel, E., 2002. Mapping of the benzodiazepine recognition site on GABA(A) receptors. Curr Top Med Chem 2, 833-839. Sigel, E., Baur, R., Furtmueller, R., Razet, R., Dodd, R.H., Sieghart, W., 2001. Differential cross talk of ROD compounds with the benzodiazepine binding site. Mol Pharmacol 59, 1470-1477. Sivilotti, L., Nistri, A., 1991. GABA receptor mechanisms in the central nervous system. Prog Neurobiol 36, 35-92. Smart, T.G., 1992. A novel modulatory binding site for zinc on the GABAA receptor complex in cultured rat neurones. J Physiol 447, 587-625. Smart, T.G., Moss, S.J., Xie, X., Huganir, R.L., 1991. GABAA receptors are differentially sensitive to zinc: dependence on subunit composition. Br J Pharmacol 103, 1837-1839. Smith, A.J., Alder, L., Silk, J., Adkins, C., Fletcher, A.E., Scales, T., Kerby, J., Marshall, G., Wafford, K.A., McKernan, R.M., Atack, J.R., 2001. Effect of alpha subunit on allosteric modulation of ion channel function in stably expressed human recombinant gamma-aminobutyric acid(A) receptors determined using (36)Cl ion flux. Mol Pharmacol 59, 1108-1118. Smith, M.C., Riskin, B.J., 1991. The clinical use of barbiturates in neurological disorders. Drugs 42, 365-378. Smith, S.S., Gong, Q.H., Hsu, F.C., Markowitz, R.S., ffrench-Mullen, J.M., Li, X., 1998. GABA(A) receptor alpha4 subunit suppression prevents withdrawal properties of an endogenous steroid. Nature 392, 926-930. Smith, S.S., Gong, Q.H., Li, X., Moran, M.H., Bitran, D., Frye, C.A., Hsu, F.C., 1998. Withdrawal from 3alpha-OH-5alpha-pregnan-20-One using a pseudopregnancy model alters the kinetics of hippocampal GABAA-gated current and increases the GABAA receptor alpha4 subunit in association with increased anxiety. J Neurosci 18, 5275-5284. Soghomonian, J.J., Martin, D.L., 1998. Two isoforms of glutamate decarboxylase: why? Trends Pharmacol Sci 19, 500-505. Speroff, L., 2005. Alternative therapies for postmenopausal women. Int J Fertil Womens Med 50, 101- 114. Spivak, C.E., 1994. Desensitization and noncompetitive blockade of GABAA receptors in ventral midbrain neurons by a neurosteroid dehydroepiandrosterone sulfate. Synapse 16, 113-122. Stell, B.M., Brickley, S.G., Tang, C.Y., Farrant, M., Mody, I., 2003. Neuroactive steroids reduce neuronal excitability by selectively enhancing tonic inhibition mediated by delta subunit-containing GABAA receptors. Proc Natl Acad Sci U S A 100, 14439-14444. Stell, B.M., Mody, I., 2002. Receptors with different affinities mediate phasic and tonic GABA(A) conductances in hippocampal neurons. J Neurosci 22, RC223. Stoffel-Wagner, B., 2003. Neurosteroid biosynthesis in the human brain and its clinical implications. Ann N Y Acad Sci 1007, 64-78. Stromberg, J., Backstrom, T., Lundgren, P., 2005. Rapid non-genomic effect of metabolites and neurosteroids on the gamma-aminobutyric acid-A receptor. Eur J Neurosci 21, 2083- 2088. Stromberg, J., Haage, D., Taube, M., Backstrom, T., Lundgren, P., 2006. Neurosteroid modulation of allopregnanolone and GABA effect on the GABA-A receptor. Neuroscience 143, 73-81. Study, R.E., Barker, J.L., 1981. Diazepam and (--)-pentobarbital: fluctuation analysis reveals different mechanisms for potentiation of gamma-aminobutyric acid responses in cultured central neurons. Proc Natl Acad Sci U S A 78, 7180-7184. Stuhmer, W., Methfessel, C., Sakmann, B., Noda, M., Numa, S., 1987. Patch clamp characterization of sodium channels expressed from rat brain cDNA. Eur Biophys J 14, 131-138. Sunderland, T., Merril, C.R., Harrington, M.G., Lawlor, B.A., Molchan, S.E., Martinez, R., Murphy, D.L., 1989. Reduced plasma dehydroepiandrosterone concentrations in Alzheimer's disease. Lancet 2, 570. Sundstro, I.I., Backstrom, T., 1999. Citalopram increases pregnanolone sensitivity in patients with premenstrual dysphoric disorder. Acta Physiol Scand 167, A6-A7.

69 Sundstrom, I., Andersson, A., Nyberg, S., Ashbrook, D., Purdy, R.H., Backstrom, T., 1998. Patients with premenstrual syndrome have a different sensitivity to a neuroactive steroid during the menstrual cycle compared to control subjects. Neuroendocrinology 67, 126-138. Sundstrom Poromaa, I., Smith, S., Gulinello, M., 2003. GABA receptors, progesterone and premenstrual dysphoric disorder. Arch Womens Ment Health 6, 23-41. Taketo, M., Yoshioka, T., 2000. Developmental change of GABA(A) receptor-mediated current in rat hippocampus. Neuroscience 96, 507-514. Thompson, S.A., Bonnert, T.P., Cagetti, E., Whiting, P.J., Wafford, K.A., 2002. Overexpression of the GABA(A) receptor epsilon subunit results in insensitivity to anaesthetics. Neuropharmacology 43, 662- 668. Thompson, S.A., Whiting, P.J., Wafford, K.A., 1996. Barbiturate interactions at the human GABAA receptor: dependence on receptor subunit combination. Br J Pharmacol 117, 521-527. Torres, J.M., Ruiz, E., Ortega, E., 2001. Effects of CRH and ACTH administration on plasma and brain neurosteroid levels. Neurochem Res 26, 555-558. Tossman, U., Jonsson, G., Ungerstedt, U., 1986. Regional distribution and extracellular levels of amino acids in rat central nervous system. Acta Physiol Scand 127, 533-545. Turkmen, S., Lofgren, M., Birzniece, V., Backstrom, T., Johansson, I.M., 2006. Tolerance development to Morris water maze test impairments induced by acute allopregnanolone. Neuroscience 139, 651-659. Turkmen, S., Lundgren, P., Birzniece, V., Zingmark, E., Backstrom, T., Johansson, I.M., 2004. 3beta- 20beta-dihydroxy-5alpha-pregnane (UC1011) antagonism of the GABA potentiation and the learning impairment induced in rats by allopregnanolone. Eur J Neurosci 20, 1604-1612. Turner, D.M., Ransom, R.W., Yang, J.S., Olsen, R.W., 1989. Steroid anesthetics and naturally occurring analogs modulate the gamma-aminobutyric acid receptor complex at a site distinct from barbiturates. J Pharmacol Exp Ther 248, 960-966. Tyndale, R.F., Bhave, S.V., Hoffmann, E., Hoffmann, P.L., Tabakoff, B., Tobin, A.J., Olsen, R.W., 1997. Pentobarbital decreases the gamma-aminobutyric acidA receptor subunit gamma-2 long/short mRNA ratio by a mechanism distinct from receptor occupation. J Pharmacol Exp Ther 283, 350-357. Tyndale, R.F., Olsen, R.W., Tobin, A.J., 1994. GABAA receptors. CRC Handbook of receptors and channels. CRC press,FL, 265-290. Ueno, S., Harris, R.A., Messing, R.O., Sanchez-Perez, A.M., Hodge, C.W., McMahon, T., Wang, D., Mehmert, K.K., Kelley, S.P., Haywood, A., Olive, M.F., Buck, K.J., Hood, H.M., Blednov, Y., Findlay, G., Mascia, M.P., 2001. Alcohol actions on GABA(A) receptors: from protein structure to mouse behavior. Alcohol Clin Exp Res 25, 76S-81S. Upasani, R.B., Yang, K.C., Acosta-Burruel, M., Konkoy, C.S., McLellan, J.A., Woodward, R.M., Lan, N.C., Carter, R.B., Hawkinson, J.E., 1997. 3 alpha-Hydroxy-3 beta-(phenylethynyl)-5 beta-pregnan-20- ones: synthesis and pharmacological activity of neuroactive steroids with high affinity for GABAA receptors. J Med Chem 40, 73-84. Valera, S., Ballivet, M., Bertrand, D., 1992. Progesterone modulates a neuronal nicotinic acetylcholine receptor. Proc Natl Acad Sci U S A 89, 9949-9953. van Rijnsoever, C., Tauber, M., Choulli, M.K., Keist, R., Rudolph, U., Mohler, H., Fritschy, J.M., Crestani, F., 2004. Requirement of alpha5-GABAA receptors for the development of tolerance to the sedative action of diazepam in mice. J Neurosci 24, 6785-6790. Vanover, K.E., Edgar, D.M., Seidel, W.F., Hogenkamp, D.J., Fick, D.B., Lan, N.C., Gee, K.W., Carter, R.B., 1999. Response-rate suppression in operant paradigm as predictor of soporific potency in rats and identification of three novel sedative-hypnotic neuroactive steroids. J Pharmacol Exp Ther 291, 1317- 1323. Vanover, K.E., Hogenkamp, D.J., Lan, N.C., Gee, K.W., Carter, R.B., 2001. Behavioral characterization of Co 134444 (3alpha-hydroxy-21-(1'-imidazolyl)-3beta-methoxymethyl-5alpha- pregnan-20-one), a novel sedative-hypnotic neuroactive steroid. Psychopharmacology (Berl) 155, 285- 291. Vanover, K.E., Rosenzweig-Lipson, S., Hawkinson, J.E., Lan, N.C., Belluzzi, J.D., Stein, L., Barrett, J.E., Wood, P.L., Carter, R.B., 2000. Characterization of the anxiolytic properties of a novel

70 neuroactive steroid, Co 2-6749 (GMA-839; WAY-141839; 3alpha, 21-dihydroxy-3beta- trifluoromethyl-19-nor-5beta-pregnan-20-one), a selective modulator of gamma-aminobutyric acid(A) receptors. J Pharmacol Exp Ther 295, 337-345. Verdoorn, T.A., Draguhn, A., Ymer, S., Seeburg, P.H., Sakmann, B., 1990. Functional properties of recombinant rat GABAA receptors depend upon subunit composition. Neuron 4, 919-928. Vicini, S., Losi, G., Homanics, G.E., 2002. GABA(A) receptor delta subunit deletion prevents neurosteroid modulation of inhibitory synaptic currents in cerebellar neurons. Neuropharmacology 43, 646-650. Wafford, K.A., Bain, C.J., Whiting, P.J., Kemp, J.A., 1993. Functional comparison of the role of gamma subunits in recombinant human gamma-aminobutyric acidA/benzodiazepine receptors. Mol Pharmacol 44, 437-442. Wafford, K.A., Ebert, B., 2006. Gaboxadol--a new awakening in sleep. Curr Opin Pharmacol 6, 30-36. Wafford, K.A., Macaulay, A.J., Fradley, R., O'Meara, G.F., Reynolds, D.S., Rosahl, T.W., 2004. Differentiating the role of gamma-aminobutyric acid type A (GABAA) receptor subtypes. Biochem Soc Trans 32, 553-556. Wafford, K.A., Thompson, S.A., Thomas, D., Sikela, J., Wilcox, A.S., Whiting, P.J., 1996. Functional characterization of human gamma-aminobutyric acidA receptors containing the alpha 4 subunit. Mol Pharmacol 50, 670-678. Wallner, M., Hanchar, H.J., Olsen, R.W., 2003. Ethanol enhances alpha 4 beta 3 delta and alpha 6 beta 3 delta gamma-aminobutyric acid type A receptors at low concentrations known to affect humans. Proc Natl Acad Sci U S A 100, 15218-15223. Wang, M., Backstrom, T., Sundstrom, I., Wahlstrom, G., Olsson, T., Zhu, D., Johansson, I.M., Bjorn, I., Bixo, M., 2001. Neuroactive steroids and central nervous system disorders. Int Rev Neurobiol 46, 421- 459. Wang, M., He, Y., Eisenman, L.N., Fields, C., Zeng, C.M., Mathews, J., Benz, A., Fu, T., Zorumski, E., Steinbach, J.H., Covey, D.F., Zorumski, C.F., Mennerick, S., 2002. 3beta -hydroxypregnane steroids are pregnenolone sulfate-like GABA(A) receptor antagonists. J Neurosci 22, 3366-3375. Wang, M., Seippel, L., Purdy, R.H., Backstrom, T., 1996. Relationship between symptom severity and steroid variation in women with premenstrual syndrome: study on serum pregnenolone, pregnenolone sulfate, 5 alpha-pregnane-3,20-dione and 3 alpha-hydroxy-5 alpha-pregnan-20-one. J Clin Endocrinol Metab 81, 1076-1082. Wang, M.D., Backstrom, T., Landgren, S., 2000. The inhibitory effects of allopregnanolone and pregnanolone on the population spike, evoked in the rat hippocampal CA1 stratum pyramidale in vitro, can be blocked selectively by epiallopregnanolone. Acta Physiol Scand 169, 333-341. Wardell, B., Marik, P.S., Piper, D., Rutar, T., Jorgensen, E.M., Bamber, B.A., 2006. Residues in the first transmembrane domain of the Caenorhabditis elegans GABA(A) receptor confer sensitivity to the neurosteroid pregnenolone sulfate. Br J Pharmacol 148, 162-172. White, G., Gurley, D., 1995. Benzodiazepine site inverse agonists can selectively inhibit subtypes of the GABAA receptor. Neuroreport 6, 1313-1316. Whitfield, G.K., Jurutka, P.W., Haussler, C.A., Haussler, M.R., 1999. Steroid hormone receptors: evolution, ligands, and molecular basis of biologic function. J Cell Biochem Suppl 32-33, 110-122. Whiting, P., McKernan, R.M., Iversen, L.L., 1990. Another mechanism for creating diversity in gamma-aminobutyrate type A receptors: RNA splicing directs expression of two forms of gamma 2 phosphorylation site. Proc Natl Acad Sci U S A 87, 9966-9970. Whiting, P.J., 2003. GABA-A receptor subtypes in the brain: a paradigm for CNS drug discovery? Drug Discov Today 8, 445-450. Whittemore, E.R., Yang, W., Drewe, J.A., Woodward, R.M., 1996. Pharmacology of the human gamma-aminobutyric acidA receptor alpha 4 subunit expressed in Xenopus laevis oocytes. Mol Pharmacol 50, 1364-1375. Wieland, S., Belluzzi, J.D., Stein, L., Lan, N.C., 1995. Comparative behavioral characterization of the neuroactive steroids 3 alpha-OH,5 alpha-pregnan-20-one and 3 alpha-OH,5 beta-pregnan-20-one in rodents. Psychopharmacology (Berl) 118, 65-71.

71 Wihlback, A.-C., Sundstrom-Poromaa, I., Backstrom, T., 2006. Action by and sensitivity to neuroactive steroids in menstrual cycle related CNS disorders. Psychopharmacology (Berl) 186, 388-401. Wingrove, P.B., Safo, P., Wheat, L., Thompson, S.A., Wafford, K.A., Whiting, P.J., 2002. Mechanism of alpha-subunit selectivity of benzodiazepine pharmacology at gamma-aminobutyric acid type A receptors. Eur J Pharmacol 437, 31-39. Wingrove, P.B., Thompson, S.A., Wafford, K.A., Whiting, P.J., 1997. Key amino acids in the gamma subunit of the gamma-aminobutyric acidA receptor that determine ligand binding and modulation at the benzodiazepine site. Mol Pharmacol 52, 874-881. Wingrove, P.B., Wafford, K.A., Bain, C., Whiting, P.J., 1994. The modulatory action of at the gamma-aminobutyric acid type A receptor is determined by a single amino acid in the beta 2 and beta 3 subunit. Proc Natl Acad Sci U S A 91, 4569-4573. Wohlfarth, K.M., Bianchi, M.T., Macdonald, R.L., 2002. Enhanced neurosteroid potentiation of ternary GABA(A) receptors containing the delta subunit. J Neurosci 22, 1541-1549. Wolkowitz, O.M., Reus, V.I., Roberts, E., Manfredi, F., Chan, T., Ormiston, S., Johnson, R., Canick, J., Brizendine, L., Weingartner, H., 1995. and cognition-enhancing effects of DHEA in major depression. Ann N Y Acad Sci 774, 337-339. Wu, F.S., Gibbs, T.T., Farb, D.H., 1991. Pregnenolone sulfate: a positive allosteric modulator at the N- methyl-D-aspartate receptor. Mol Pharmacol 40, 333-336. Xue, B.G., Whittemore, E.R., Park, C.H., Woodward, R.M., Lan, N.C., Gee, K.W., 1997. Partial agonism by 3alpha,21-dihydroxy-5beta-pregnan-20-one at the gamma-aminobutyric acidA receptor neurosteroid site. J Pharmacol Exp Ther 281, 1095-1101. Yamada, J., Okabe, A., Toyoda, H., Kilb, W., Luhmann, H.J., Fukuda, A., 2004. Cl- uptake promoting depolarizing GABA actions in immature rat neocortical neurones is mediated by NKCC1. J Physiol 557, 829-841. Yang, J., Kumanovic, S., Xia, H., Fick, D.B., Nguyen, P.S., Espitia, S.A., Hawkinson, J.E., Vanover, K.E., Carter, R.B., Lan, N., Hogenkamp, D.J., 1999. Neuroactive steroids: synthesis and SAR of 21-(N- heteroaryl)-3α-hydroxy-3β-methyl-5α-pregnan-20-ones. Book of abstractss, 217th ACS national meeting, Anaheim, CA, United States March 21-25. Yee, B.K., Hauser, J., Dolgov, V.V., Keist, R., Mohler, H., Rudolph, U., Feldon, J., 2004. GABA receptors containing the alpha5 subunit mediate the trace effect in aversive and appetitive conditioning and extinction of conditioned fear. Eur J Neurosci 20, 1928-1936. Yeung, J.Y., Canning, K.J., Zhu, G., Pennefather, P., MacDonald, J.F., Orser, B.A., 2003. Tonically activated GABAA receptors in hippocampal neurons are high-affinity, low-conductance sensors for extracellular GABA. Mol Pharmacol 63, 2-8. Ymer, S., Draguhn, A., Wisden, W., Werner, P., Keinanen, K., Schofield, P.R., Sprengel, R., Pritchett, D.B., Seeburg, P.H., 1990. Structural and functional characterization of the gamma 1 subunit of GABAA/benzodiazepine receptors. Embo J 9, 3261-3267. Zhu, D., Wang, M.D., Backstrom, T., Wahlstrom, G., 2001. Evaluation and comparison of the pharmacokinetic and pharmacodynamic properties of allopregnanolone and pregnanolone at induction of anaesthesia in the male rat. Br J Anaesth 86, 403-412. Zhu, W.J., Vicini, S., 1997. Neurosteroid prolongs GABAA channel deactivation by altering kinetics of desensitized states. J Neurosci 17, 4022-4031. Zorumski, C.F., Mennerick, S., Isenberg, K.E., Covey, D.F., 2000. Potential clinical uses of neuroactive steroids. Curr Opin Investig Drugs 1, 360-369. Zwain, I.H., Yen, S.S., 1999. Neurosteroidogenesis in astrocytes, oligodendrocytes, and neurons of cerebral cortex of rat brain. Endocrinology 140, 3843-3852.

72